Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2018 Jan 1.
Published in final edited form as: Adv Exp Med Biol. 2017;964:133–152. doi: 10.1007/978-3-319-50174-1_10

Sigma-1 Receptors and Neurodegenerative Diseases: Towards a Hypothesis of Sigma-1 Receptors as Amplifiers of Neurodegeneration and Neuroprotection

Linda Nguyen 1,2, Brandon P Lucke-Wolds 3, Shona Mookerjee 4, Nidhi Kaushal 5, Rae R Matsumoto 6,7,
PMCID: PMC5500918  NIHMSID: NIHMS871113  PMID: 28315269

Abstract

Sigma-1 receptors are molecular chaperones that may act as pathological mediators and targets for novel therapeutic applications in neurodegenerative diseases. Accumulating evidence indicates that sigma-1 ligands can either directly or indirectly modulate multiple neurodegenerative processes, including excitotoxicity, calcium dysregulation, mitochondrial and endoplasmic reticulum dysfunction, inflammation, and astrogliosis. In addition, sigma-1 ligands may act as disease-modifying agents in the treatment for central nervous system (CNS) diseases by promoting the activity of neurotrophic factors and neural plasticity. Here, we summarize their neuroprotective and neurorestorative effects in different animal models of acute brain injury and chronic neurodegenerative diseases, and highlight their potential role in mitigating disease. Notably, current data suggest that sigma-1 receptor dysfunction worsens disease progression, whereas enhancement amplifies pre-existing functional mechanisms of neuroprotection and/or restoration to slow disease progression. Collectively, the data support a model of the sigma-1 receptor as an amplifier of intracellular signaling, and suggest future clinical applications of sigma-1 ligands as part of multi-therapy approaches to treat neurodegenerative diseases.

Keywords: Neurorestoration, Stroke, Parkinson’s disease, Alzheimer’s disease, Amyotrophic lateral sclerosis

10.1 Introduction

Neurodegeneration is characterized by the loss of neuronal integrity, in both structure and function, and can result from acute injury or chronic disease progression. Neurodegenerative diseases are a major cause of morbidity and mortality among the aging population worldwide. The World Alzheimer Report, for example, estimates 46.8 million people worldwide are living with dementia as of 2015 [1]. Based on current trends, this number is projected to almost double about every 20 years, reaching 74.7 million in 2030 and 131.5 million in 2050 [1, 2], making efforts to understand and treat these conditions crucial to maintaining the health of an increasingly large demographic. While current treatments for neurodegenerative conditions such as Alzheimer’s disease (AD), Parkinson’s disease (PD), Huntington’s disease (HD), amyotrophic lateral sclerosis (ALS), and stroke can temporarily alleviate some symptoms and improve quality of life, they are generally ineffective at slowing or stopping disease progression.

Sigma receptors are increasingly recognized targets for novel therapeutic intervention in neurodegenerative conditions [3]. These proteins are implicated in diverse neural mechanisms, including the modulation of cell survival and function, calcium signaling, neurotransmitter release, inflammation, and synaptogenesis [46]. The two established subtypes of sigma receptors, sigma-1 and sigma-2, are highly expressed in the central nervous system (CNS), and are distinguishable pharmacologically, functionally, and by molecular size [7].

The present chapter focuses on putative neuroprotective roles conferred by sigma-1 receptor activity as this receptor is the better characterized of the two subtypes. If both subtypes may be involved in an effect described here, we refer to sigma receptors in general, rather than specifying the subtype. We begin with a brief overview of the neuroprotective and restorative effects of sigma-1 ligands in various animal models of neurodegenerative diseases. “Neuroprotection” and “neurorestoration” are terms that can be interpreted in a variety of ways. In the context of this review, neuroprotection is any sequence of events that interrupts or slows the sequence of injurious biochemical and molecular events that, if left unchecked, would likely lead to cell damage and/or loss. Neurorestoration is the regeneration of functional tissue, which is impacted by the capacity of surviving cells to adapt after injury and of new cells (through neurogenesis and/or recruitment of glial cells to damaged areas) to support repair. This is followed by a summary of how sigma-1 ligands may confer their therapeutic effects by modulating mechanisms that are common across a wide array of neurodegenerative conditions, including excitotoxicity, Ca+2 dysregulation, mitochondrial and endoplasmic reticulum (ER) dysfunction, neuroinflammation, and reactive gliosis. In addition, sigma-1 ligands may promote neurorestorative processes to enhance the structure and function of neurons that become compromised in disease, or to stimulate the influx of new cells to assist in repairing damage to the nervous system. Finally, we discuss the model of the sigma-1 receptor as an “amplifier” of intracellular signaling, the resulting ways that sigma-1 receptors may be involved in disease and therefore exploited therapeutically, and the potential application of sigma-1 ligands as part of combined therapeutic approaches in future clinical studies of neurodegenerative diseases.

10.2 Sigma-1 Receptor Ligands in Animal Models of Neurodegenerative Diseases

In general, in model systems of neurodegenerative diseases, deficits in sigma-1 receptor level or activity are associated with neurodegeneration, while sigma-1 receptor activation or overexpression are associated with neuroprotection. Consistent with this, most reports of beneficial effect come from studies of sigma-1 agonists, with these effects generally showing sensitivity to sigma-1 antagonism. Below, we highlight the neuroprotective effects of sigma-1 receptors in several animal models of neurodegenerative disease. In our recent review, we highlight additional models where targeting sigma-1 receptors alters neurodegenerative disease processes [3].

10.2.1 Stroke

Acute brain injury following cerebral ischemia (stroke) and trauma can lead to long-term neurological and psychiatric deficits. As the primary insult (e.g., direct mechanical damage) cannot be therapeutically influenced, the goal of treatment is to limit secondary injury processes. Following cerebral ischemia, both necrotic and apoptotic cell death can be induced through complex interactions of pathological processes, including excitotoxicity and inflammation [8, 9]. Neuroprotective and neurorestorative effects of sigma-1 agonists (e.g., decreasing cell death, protecting against tissue damage, and increasing synaptic protein expression) have been shown in multiple animal models of stroke, including mouse [10], rat [1116], gerbil [17] and cat [18]. In rat models of stroke, for example, decreased infarct volume as well as enhanced neuronal survival were observed following acute treatment with a sigma agonist 24 h after the onset of ischemia [14, 15]. In addition, functional recovery with or without changes in infarct volume was observed when sigma agonists were administered as late as 2 days post-stroke [15, 19]. The potential to treat at extended times following the initial embolic injury warrants further investigation, as the only available post-stroke treatment approved for use in humans is thrombolytics, which is limited to 4 h post-stroke due to the risk of hemorrhagic transformation (i.e., conversion of an ischemic stroke to a hemorrhagic one following reperfusion) [20]. Of note, Ruscher and colleagues demonstrated that treatment of rats subjected to permanent or transient middle cerebral artery occlusion (MCAO) with the selective sigma-1 agonist SA4503 (1-[2-(3,4-dimethoxyphenyl) ethyl]-4-(3-phenylpropyl) piperazine) starting 2 days after injury conferred significantly better recovery rates of sensorimotor function compared with the vehicle group [19]. The significant improvement of neurological function following MCAO was associated with increased levels of the synaptic proteins neurabin and neurexin in the peri-infarct area [19]. This improvement was sustained 2 weeks after discontinuation of SA4503 [19]. These results suggest that stimulation of sigma-1 receptors promote neural adaptations (e.g., increases in synaptic proteins and potentially synaptic connections) to facilitate recovery following MCAO [19].

10.2.2 Other Acute CNS Injury

Additional beneficial effects of sigma agonists have been reported in other models of acute CNS injury. Following spinal root avulsion in adult rats, administration of the sigma-1 agonist PRE084 (2-(4-morpholinethyl)1-phenylcyclohexanecarboxylate) promoted motor neuron survival [21]. Co-administration of the sigma-1 antagonist BD1063 (1-[2-(3,4-dichlorophenyl) ethyl]-4-methylpiperazine) blocked this effect [21]. In another study, the sigma agonist PPBP (4-phenyl-1-(4-phenylbutyl) piperidine) improved neurological function and reduced striatal cell death when administered after global hypoxia-ischemia (induced by asphyxic cardiac arrest followed by resuscitation) in newborn piglets [22]. An additional study showed that the sigma-1 agonist PRE084 reduced cortical lesion size and cell death following excitotoxic perinatal brain injury in newborn mice [23]. However, confirmation of a sigma-1-mediated mechanism using an antagonist was not tested in the two aforementioned studies involving neonates.

10.2.3 Amyotrophic Lateral Sclerosis

In a superoxide dismutase 1 (SOD1)-G93A mouse model of amyotrophic lateral sclerosis (ALS), daily administration of the selective sigma-1 agonist PRE084 from 8 to 16 weeks of age improved spinal motor neuron function and survival, demonstrated by the preservation of neuromuscular connections and motor neuron number in the spinal cord, maintenance of muscle action potential amplitudes, and improvement in locomotor performance [24]. This attenuation of the disease state was associated with an increase in survival time in PRE084-treated mice compared to controls [24]. In contrast, genetic ablation of sigma-1 receptors accelerated the appearance of motor deficits as well as decreased longevity in the SOD1-G93A mouse model [25].

Sigma-1 agonists may also be effective in cases of SOD-1-independent mechanisms of ALS. Chronic treatment with PRE084 improved motor neuron survival and locomotor performance in the wobbler mouse, which is a model of motor neuron degeneration [26].

10.2.4 Parkinson’s Disease

In an intrastriatal 6-hydroxydopamine (6-OHDA) lesion model of Parkinson’s disease (PD), mice were treated daily with PRE084 for 5 weeks, starting on the same day as the lesion induction [27]. PRE084 gradually and significantly improved spontaneous forelimb use, along with a partial recovery of dopamine levels and increased dopaminergic fiber densities, compared to saline-treated animals [27]. PRE084 treatment also upregulated neurotrophic factor protein levels and increased activation of their downstream effector pathways [27], further suggesting that sigma-1 receptor activation contributes to the restoration of synaptic connectivity and functional recovery in neurodegeneration disease models.

10.2.5 Alzheimer’s Disease

In an amyloid beta (25–35) peptide-induced mouse model of Alzheimer’s disease (AD), selective and non-selective sigma-1 agonists improved both molecular and behavioral markers of neurodegeneration [2, 28, 29]. The selective sigma-1 agonist PRE084 and the non-selective sigma-1 agonists donepezil or AVANEX2–73 mitigated spatial working memory deficits in spontaneous alternation tests [2, 28]. They also attenuated contextual long-term memory in the step-through passive avoidance procedure [2, 28]. These effects were mediated, at least in part, by sigma-1 receptors, demonstrated by their sensitivity to the sigma-1 antagonist BD1047 (N-[2-(3, 4-dichlorophenyl) ethyl]-N-methyl-2-(dimethylamino) ethylamine) [2, 28]. In addition, treatment with these sigma-1 agonists decreased amyloid beta-induced lipid peroxidation in the hippocampus, consistent with a role in decreasing oxidative damage; these protective effects were also attenuated by BD1047 [2, 28]. In amyloid beta (25–35)-treated mice exhibiting cognitive deficits, PRE084 or igmesine, another selective sigma-1 agonist, showed greater antidepressant efficacy compared to non-amyloid beta-treated animals [6]. This enhanced efficacy was not seen with the classic antidepressants desipramine or fluoxetine, suggesting that selective sigma-1 receptor agonists are promising alternatives for alleviating the depressive symptoms in AD patients.

10.2.6 Possible Therapeutic Effects of Sigma-1 Antagonism

There are a few studies that suggest a potential benefit of sigma-1 antagonism to promote neuroprotection, though the specific effects of sigma-1 suppression are unclear. The putative sigma-1 antagonist haloperidol, for instance, reduced infarct volume in a rat model of MCAO [30]. When compared to eight other butyrophenone compounds in an in vitro assay of glutamate-induced oxidative stress, the authors found a significant positive correlation between haloperidol’s protective potency (i.e., nanomolar vs. micromolar concentration required to increase cell survival) and affinity for sigma-1 receptors [30]; however, haloperidol also has similar nanomolar affinity to other targets, including dopamine, serotonin (5-HT), and alpha adrenergic receptors, making it difficult to attribute its primary effect to sigma-1 antagonism in these models. More selective sigma antagonists have been shown to reduce methamphetamine (METH)-induced neurotoxicity [31] and alleviate neuropathic pain [32]. In wildtype mice, knockout of sigma-1 receptors prevented subchronic administration of the dopaminergic neurotoxin 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP) from causing the motor and histochemical deficits characteristic of PD [33]. This protective effect, however, was not observed in sigma-1 knockout mice [33], suggesting the importance of these receptors in the etiology of the disease. While these and other studies leave open the possibility that sigma-1 antagonism may be beneficial in certain conditions, there is much stronger and more direct evidence that sigma-1 activation is neuroprotective, and therefore that sigma-1 agonist-based therapeutics are more likely to protect against neurodegeneration than antagonists.

10.3 Sigma-1 Receptor Mediated Mechanisms of Neuroprotection

While neurodegenerative diseases are a heterogeneous group of illnesses with distinct clinical phenotypes and diverse etiologies, emerging evidence suggests that they share important pathogenic mechanisms, including excitotoxicity [20, 34, 35], Ca2+ dysregulation [36, 37], mitochondrial and ER dysfunction [3841], inflammation [42, 43], and in some cases, astrogliosis [44]. In addition, neurotrophic factors and neural plasticity have been found to be important targets for disease-modifying treatments for CNS diseases [4548]. In this section, we focus on the ways in which sigma-1 receptor activity may modulate these mechanisms to elicit neuroprotection.

10.3.1 Glutamate Excitotoxicity

Excitotoxicity occurs when high levels of glutamate cause persistent activation of N-methyl-D-aspartate (NMDA) receptors, allowing an influx of Ca2+ that can activate downstream mechanisms of programmed cell death, including the activation of calpains, proteases, protein kinases, nitric oxide synthase (NOS) and the mitochondrial permeability transition pore [34, 49]. Excitotoxicity has been observed in multiple neurodegenerative disease states, including ALS, AD, PD, stroke and METH toxicity [20, 26, 35, 50, 51]. Through the modulation of glutamate and its receptors, sigma ligands have been reported to be neuroprotective against excitotoxicity in retinal ganglion cells (RGCs), primary neuronal cultures, and ischemic stroke models [23, 5257].

The mechanisms by which sigma ligands modulate excitotoxic glutamate release are poorly understood. However, studies to date implicate multiple mechanisms. In a chronic restraint stress model of depression, for example, stimulation of sigma-1 receptors enhanced glutamate release by increasing presynaptic cytoplasmic release of Ca2+ from ER stores [58]. Sigma-1 agonists also inhibited the release of glutamate evoked by a K+ channel blocker in cortical nerve endings, in a sigma-1 antagonist-sensitive manner [59]. In addition, treatment with sigma-1 agonists has led to decreased Ca2+ entry through presynaptic voltage-dependent Ca2+ channels and the suppression of protein kinase C (PKC) signaling cascades, resulting in decreased glutamate release from nerve terminals in the rat cerebral cortex [59].

In addition to influencing glutamate release, sigma-1 receptor activity is implicated in the neuronal responses to NMDA receptor stimulation, both directly, through interactions with specific subunits of the NMDA receptor [60, 61] and indirectly, through the modulation of other ion channels [62]. Sigma-1 receptors have been shown to bind to the cytosolic C-terminal region of the NMDA receptor NR1 subunit in recombinant cells, which can be inhibited by sigma-1 antagonists [63]. Activation of sigma-1 receptors can also increase the interaction between sigma-1 receptors and NR2 subunits of NMDA receptors. This happens concurrently with increased translocation to the cell surface and results in an increase in NMDA receptor availability at the plasma membrane [64]. The authors hypothesized that the relationship between sigma-1 receptors and NR2 subunits is therefore an indirect one, involving direct interactions between sigma-1 receptors and NR1 subunits that are part of the same tetrameric NMDA receptor complex as the NR2 subunits being probed [64]. In another study, activation of sigma-1 receptors induced phosphorylation of NR1 subunits and subsequent potentiation of NMDA receptor function in spinal neurons by modulating PKC signaling via the alpha and epsilon isoforms of PKC [65].

Sigma-1 receptor activation can also affect the interaction of other proteins with NMDA receptors to elicit neuroprotective effects. For example, sigma-1 agonists enhanced the interaction of histidine triad nucleotide binding protein 1 (HINT1) with G-protein coupled receptors (GPCRs) and in turn stimulated GPCR-NMDA interactions, promoting protective effects against excitotoxicity [66]. Downstream of influencing NMDA receptor function and/or activity, sigma-1 agonists have been shown to be neuroprotective by increasing brain-derived neurotrophic factor (BDNF) levels in an ischemia/reperfusion vascular dementia model [67]. This appeared to be mediated through NR2A-CAMKIV (calcium/calmodulin-dependent protein kinase type IV)-TORC1 (transducer of regulated cyclic adenosine monophosphate (cAMP) responsive element-binding protein (CREB) activity) pathways [67].

In addition to NMDA receptors, sigma receptors may regulate (directly or indirectly) other glutamatergic targets including kainate and alpha-amino-3-hydroxy-5-methyl-4-isoxazolepropionic (AMPA) glutamate receptors to confer neuroprotective effects. For example, sigma-1 agonists attenuated kainate receptor-induced hippocampal neurotoxicity and seizures by acting downstream and decreasing c-fos/c-jun expression and activator protein (AP)-1 DNA-binding activity [68, 69]. Sigma-1 agonism also afforded neuroprotection by reducing the expression of AMPA receptors in cultured cortical neurons, possibly via decreasing activation of mitogen-activated protein kinase (MAPK)/extracellular signal-regulated kinase (ERK) signaling [70], a critical pathway for the maintenance of ionotropic glutamate receptors [71].

10.3.2 Ca2+ Dysregulation

High and persistent Ca2+ release may contribute to neurotoxicity and cell death. In addition to Ca2+ flux through NMDA receptors, there are several additional means by which Ca2+ levels can be increased to toxic levels in neurons, including exit from intracellular ER and mitochondria stores, influx through voltage-dependent plasma membrane Ca2+ channels, Na2+/Ca2+ exchangers, and acid-sensing ion channels (ASICs) [49].

Sigma-1 agonists have been shown to regulate intracellular Ca2+ levels and prevent the increased expression of pro-apoptotic genes and caspases in RGCs [72], as well as in rat cortical neurons with prolonged exposure to amyloid beta peptide [73]. These molecular effects correspond with phenotypic improvements to memory impairments in animal models [29].

In both physiological and pathophysiological conditions, sigma-1 receptors appear to function as chaperones and Ca2+ sensors [5, 7477]. At the ER mitochondrial-associated membrane (MAM), sigma-1 receptors play an important role in regulating Ca2+ levels via inositol trisphosphate (IP3) receptors and maintaining intracellular Ca2+ homeostasis [76].

In addition to modulation of intracellular sources of Ca2+, sigma-1 receptors can alter the behavior of plasma membrane ion channels, thereby altering Ca2+ uptake into the cell. Sigma-1 agonists have been shown to mediate the elevated intracellular Ca2+ levels caused by activation of ASIC-1a during stroke-induced ischemia [78]. Among the Ca2+-associated downstream signaling pathways, sigma-1 agonism reduced the activation of the MAPK/ERK pathway, affording neuroprotection [70]. In rat primary ganglion cells, the sigma-1 agonist (+)-SKF10047 ((2S,6S,11S)-1,2,3,4,5,6-hexahydro-6,11-dimethyl-3-(2-propenyl)-2,6-methano-3-benzazocin-8-ol) inhibited potassium chloride (KCl)-induced Ca2+ influx through L-type Ca2+ channels, which was reversed by the sigma-1 antagonist, BD1047. This inhibition involved direct interactions between L-type Ca2+ channels and sigma-1 receptors [79]. In addition to their effects on neurons, sigma-1 ligands can suppress microglial activation through Ca2+-dependent mechanisms, decreasing the release of inflammatory cytokines [80].

10.3.3 ER Stress

Neurodegenerative conditions associated with ER stress include METH toxicity, HD, AD, ALS and PD [38, 40, 81]. One result of ER dysfunction is the accumulation of unfolded or misfolded proteins within the ER lumen. This accumulation activates the unfolded protein response (UPR), which occurs through three major signaling pathways: protein kinase RNA like ER kinase (PERK), inositol requiring enzyme 1 alpha (IRE1α), and activating transcription factor 6 (ATF6). As chaperones, sigma-1 receptors participate closely in the degradation of unfolded proteins [82], and multiple studies describe sigma-1 receptor modulation of the UPR [76, 83, 84]. Moreover, the C-terminus on the sigma-1 receptor has been shown to interact with the glucose-regulated protein 78 (GRP78)/immunoglobulin heavy-chain binding protein (BiP) [85], a critical regulator of all three arms of the UPR [86]. Following the administration of the Ca2+ channel inhibitor thapsigargin or the GPT (UDP-N-acetylglucosamine-dolichol phosphate N-acetylglucosamine-1-phosphate transferase) inhibitor tunicamycin, which are frequently used to model ER stress and induce UPR in cell culture models, sigma-1 receptor expression is upregulated in response to activation of the PERK pathway [87] and more specifically, ATF4, a downstream target of PERK signaling [88]. This upregulation of sigma-1 receptor expression found in HEK293 (human embryonic kidney) and Neuro2a (mouse neuroblastoma) cells can repress cell death signals that accompany ER stress [87]. Consistent with this, overexpression of the sigma-1 receptor decreased the activation of PERK and ATF6 and increased cell survival in Chinese hamster ovary (CHO) cells, whereas knockdown of sigma-1 receptors destabilized the conformation of IRE1 and decreased cell survival following administration of thapsigargin [76, 84].

Not surprisingly, treatment with the selective sigma-1 agonist SA4503 mitigated ER stress and reduced cell death in the retina following light-induced damage [89]. Using the selective 5-HT reuptake inhibitor (SSRI) fluvoxamine, a potent sigma-1 agonist that exhibits stronger affinity for sigma-1 receptors than other SSRIs [90], Omi and colleagues also showed that fluvoxamine, via activation of sigma-1 receptors, upregulated sigma-1 receptor expression and inhibited cell death in Neuro2a cells exposed to tunicamycin (which disrupts protein folding and directly induces the UPR) [88]. The specificity of sigma-1 involvement was confirmed with the addition of the sigma-1 antagonist NE100 (N, N-dipropyl-2-[4-methoxy-3-(2-phenylethoxy) phenyl] ethylamine monohydrochloride), which blocked the effects of fluvoxamine [88].

10.3.4 Mitochondrial Ca2+ Uptake and Activity

In their quiescent state, sigma-1 receptors are located in the mitochondrial-associated membrane (MAM) in association with the ER chaper-one protein BiP, and under these conditions are inactive. Activation by ligand binding or various pathways of ER stress causes the dissociation of sigma-1 receptors from BiP and allow their participation in multiple downstream pathways.

Within the MAM, activated sigma-1 receptors appear to stabilize IP3 receptors by protecting them from proteasomal degradation [76] and activate them by promoting their dissociation from the ion channel chaperone protein ankyrin B 220 [91]. This promotes Ca2+-induced Ca2+ release from the ER through Ca2+-activated IP3 production, and then Ca2+ trafficking into the mitochondria. Ca2+ uptake into the mitochondrial matrix is a sensitive regulator of oxidative phosphorylation. Sub-micromolar increases in matrix Ca2+ directly activate multiple enzymes, including glycerol phosphate dehydrogenase, isocitrate dehydrogenase, and oxoglutarate dehydrogenase, and, indirectly (via dephosphorylation) pyruvate dehydrogenase, resulting in increased flux through the Krebs cycle and facilitating increased rates of oxidative phosphorylation. In this way, Ca2+ acts as an interorganellar signal to “tune” ATP supply according to the ATP demand dictated by the rest of the cell.

The sigma-1 receptor-IP3 receptor interaction may promote mitochondrial Ca2+ uptake and, ultimately, cell survival [76]. Consistent with this, Shioda and colleagues identified a truncated splice variant of the sigma-1 receptor (short form sigma-1 or sigma-1S) in the mouse hippocampus that localizes to the MAM and complexes with non-truncated sigma-1 receptors, but does not complex with IP3 receptors [92]. In Neuro2a C3100 cells, exogenous overexpression of non-truncated sigma-1 receptors enhanced ATP-or IP3-induced mitochondrial Ca2+ uptake whereas overexpression of sigma-1S decreased mitochondrial Ca2+ uptake compared to control cells [92]. Following tunicamycin-induced ER stress, the exogenous overexpression of non-truncated sigma-1 receptors protected IP3 receptor proteins from degradation and enhanced ATP production, promoting cell survival [92]. In contrast, overexpression of sigma-1S enhanced IP3 receptor degradation and decreased mitochondrial Ca2+ uptake, resulting in increased apoptosis [92]. These findings suggest that sigma-1S destabilizes IP3 receptors and diminishes IP3 receptor-driven mitochondrial Ca2+ uptake through the loss of sigma-1-IP3 receptor interactions, resulting in impaired ATP production and increased apoptosis [92]. Of note, mutations of sigma-1 receptors have been found in neurodegenerative conditions such as ALS [93, 94]. It will therefore be important to further evaluate how truncated sigma-1 receptors may interfere with normal receptor function to affect mitochondrial stability.

There is also experimental support for sigma-1 receptor-mediated maintenance of bioenergetic homeostasis. Though the effects are likely indirect, sigma-1 receptor activation is reported to preserve bioenergetic function in multiple models, supporting a neuroprotective role. The sigma agonist PPBP appears to stabilize mitochondrial membrane potential in neurons undergoing excitotoxic stress through glutamate exposure. This stabilization was associated with decreased neuronal death [95]. Another agonist, BHDP (N-benzyl-N-(2-hydroxy-3,4-dimethoxybenzyl)-piperazine), appeared to have “mitochondrial protective” effects in a liver model of ischemia/reperfusion [96].

10.3.5 Neuroinflammation

The primary mediators of neuroinflammation in the CNS are microglia, which are macrophage-derived cells residing in the CNS. Although multiple microglial phenotypes are believed to result from CNS insult, they are typically classified as M1 and/or M2 responses, similar to peripheral macrophages [42]. M1 microglia are traditionally considered pro-inflammatory and tend to be associated with damage to the CNS, while M2 microglia are anti-inflammatory and associated with neuronal repair and regrowth [42, 97]. Sigma-1 receptors, expressed in microglial and neurons, may modulate microglial activation and dampen neuroinflammation. Indeed, many studies have shown that sigma agonists may affect M1 and/or M2 responses, with most studies to date focusing on the M1 response. For example, Robson and colleagues demonstrated that neurotoxic dosing with METH preferably activated M1 microglia responses within the mouse striatum as represented by significant increases in the pan-macrophage markers, cluster of differentiation 68 (CD68) and ionized calcium binding adapter molecule 1 (IBA-1), without concurrent increases in an M2 marker, CD163 [98]. Pretreatment with the sigma ligand SN79 (6-acetyl-3-(4-(4-(4-fluorophenyl) piperazin-1-yl) butyl) benzo[d] oxazol-2(3H)-one) attenuated the increase in CD68 and IBA-1, indicating prevention of METH-induced M1 microglial activation [98]. Associated with this reduction in M1 microglia was an obviation of IL-6 and oncostatin M, showing protection against neuroinflammation [98]. In lipopolysaccharide (LPS)-stimulated murine microglial BV2 cells, the sigma-1 agonist SKF83959 (6-chloro-2,3,4,5-tetrahydro-3-methyl-1-(3-methylphenyl)-1H-3-benzazepine-7,8-diol) prevented M1 microglial activation and decreased pro-inflammatory cytokines, including tumor necrosis factor alpha, IL-1 beta, and inducible NOS [99]. The sigma agonists DTG (1,3-di-(2-tolyl)guanidine) and afobazole have also been shown to suppress microglial activation and migration and the release of inflammatory cytokines in response to not only LPS, but also other microglial activators such as ATP, uridine triphosphate and monocyte chemoattractant protein-1 [100]. In an in vivo model of traumatic brain injury, the sigma-1 agonist PRE084 has been shown to reduce IBA-1 expression following controlled cortical impact in association with reduced lesion volume and improved behavior in mice [101]. Similarly, PRE084 also reduced counts of IBA-1 positive microglial cells in a mouse model of ALS [24].

In contrast, in another study in animals with motor neuron disease, treatment with PRE084 increased the number of cells positive for the pan-macrophage marker CD68 and CD206, which is associated with M2 microglial responses [26]. Sigma ligands also improved microglial cell survival during and at least 24 h after ischemia [100], as well as after toxic exposure to amyloid beta in primary microglia cultures [102]. These data suggest that sigma receptors may modify microglial reactivity to strengthen the reparative microglia phenotype (M2) while attenuating the inflammatory response (M1). Further studies should examine both microglial types to better understand the role of sigma-1 receptors on microglia in neurodegeneration and neurorestoration.

Upon disruption of the blood brain barrier by injury, diapedesis of peripheral leukocytes into the brain may also exacerbate neuroinflammation. Sigma-1 receptors are expressed in lymphocytes, and previous studies have shown the ability of sigma-1 ligands to inhibit CD3 lymphocyte proliferation in vitro and LPS-induced release of cytokines in vivo [103, 104]. Recently, a novel synthetic, high-affinity and selective sigma-1 ligand was examined in a mouse autoimmune encephalitis model [105], which exhibited histopathological changes characterized by peripheral leukocyte infiltration into the brain along with demyelination and axonal loss [106]. The sigma-1 ligand prevented mononuclear cell accumulation and demyelination in the brain and spinal cord while also increasing the proportion of B-cell subsets and regulatory T-cells, resulting in an overall reduction of the clinical signs of experimental autoimmune encephalitis [105]. Sigma-1 receptors may therefore regulate peripheral immune cells to slow the progression of certain CNS diseases.

10.3.6 Reactive Astrogliosis

Reactive astrogliosis is characterized by the “activation” of astrocytes within the CNS, which may result in the proliferation and migration of astrocytes to areas of damage and in some cases, the formation of glial scars. Glial scar formation is hypothesized to protect surrounding neuronal tissue from further damage as a result of excess inflammation. However, the formation of glial scars also can impede repair and thereby inhibit the ability of neuronal tracts to regenerate [107]. Sigma receptors are found in astrocytes [7], and accumulating data suggests that sigma-1 receptor activity promotes repair following damage to the CNS. For example, following experimental stroke, a 30 % increase of sigma-1 receptor expression was found in astrocytes, and treatment with a sigma-1 agonist enhanced the recovery of sensorimotor function without decreasing infarct size [19]. This led the authors to suggest that the recovery-promoting action of sigma-1 receptors involved astrocytes in the peri-infarct area [19].

In an animal model of PD, sigma-1 receptor distribution was also noted in astrocytes, in addition to neurons. In the striatum, sigma-1 immunoreactivity predominated in astrocytes (vs. neurons), while in the substantia nigra, sigma-1 immunoreactivity was predominant in dopaminergic neurons (vs. astrocytes) [27]. Following 5 weeks of treatment with the sigma-1 agonist PRE084, there was a shift in distribution of sigma-1 receptors from the neuronal and astrocytic cell bodies into the processes [27]. This occurred in association with functional motor recovery. Consequently, sigma-1 agonists may be increasing astrocytic neuroprotective activity by promoting the intracellular trafficking of sigma-1 receptor proteins and potential transport of other protein partners involved in neuroprotective mechanisms to distal regions of astrocytes.

As well as increasing astrocytic activity, sigma ligands have also been shown to reduce astrocytic activity in some studies. Ajmo and colleagues showed that the sigma agonist DTG reduced astrocyte activation 24 h post-stroke [14]. Penas and colleagues also found that the selective sigma-1 agonist PRE084 reduced astrogliosis and ER stress following spinal root avulsion [108]. PRE084 similarly provided neuroprotection and reduced astrogliosis in preclinical ALS [26] and PD models [27]. Furthermore, in an animal model of METH-induced neurotoxicity, METH increased astrogliosis in the damaged striatum, an effect that could be mitigated with a novel sigma ligand; these effects involved modulation of the OSMR-STAT3 signaling pathway [109], which has been implicated in other neurotoxic conditions such as ischemic stroke or peripheral LPS injections [110].

The ability of sigma ligands to modulate the function of astrocytes, in addition to neurons and microglia, suggest that these ligands may be able to facilitate a coordinated response across cell types to achieve therapeutic outcomes. Additional studies to delineate the interplay between these cell types in the nervous system in health and disease are warranted.

10.4 Sigma-1 Receptor-Mediated Mechanisms of Neurorestoration

In addition to targeting sigma-1 receptors to mitigate neurodegenerative processes, accumulating evidence reveals the potential for sigma-1 agonists to stimulate neurorestorative processes after the CNS has been damaged. This can be achieved either by improving the structural and/or functional integrity of existing cells that have become compromised by disease, or by stimulating the incorporation of new cells into the damaged region to support repair. For example, in a mouse 6-OHDA lesion model of PD, treatment with the sigma-1 agonist PRE084 for 5 weeks after lesion induction can rescue animals from both histological and functional deficits that are normally associated with the lesions [27]. The involvement of sigma-1 receptors in the neurorestorative effects has been confirmed by the inability of PRE084 to evoke similar rescue in sigma-1 receptor knockout mice [27]. Specific neurorestorative mechanisms that can be targeted by sigma ligands are only just beginning to be investigated, and those that have been identified are summarized below.

10.4.1 Increased Growth Factor Expression or Activity

Neurotrophins and growth factors play an integral role in nervous system development, maintenance, and plasticity [111]. Aberrant levels of multiple neurotrophins and growth factors have been implicated in CNS disorders, including neurodegenerative conditions [112]. These proteins have also been proposed as targets for future therapies [113].

Glia-derived neurotrophic factor (GDNF) has long been known to be capable of rescuing neurons following CNS insult [114, 115]. Converging lines of evidence now indicate that sigma-1 receptor activation may stimulate these GDNF-dependent repair mechanisms. In a spinal root avulsion model, daily administration of the sigma-1 agonist PRE084 after axotomization of motor neurons increased motor neuron survival at day 21 post-operation [108]. This recovery was accompanied by an early increase in GDNF expression in astrocytes in the ventral horn day 3 post-operation [108], suggesting that the activation of sigma-1 receptors in glial cells led to the release of survival-promoting trophic factors. The restorative effects to behavior and histology associated with subchronic treatment of PRE084 in the mouse model of PD described above was also accompanied by an upregulation of striatal GDNF and BDNF; GDNF was additionally upregulated in the substantia nigra [27]. Since phosphorylated ERK1/2 and protein kinase B were also increased under these conditions, the data suggest that downstream signaling pathways associated with these trophic factors were activated by PRE084 to promote recovery [27].

In addition, in a mouse model of ALS, subchronic treatment with PRE084 upon symptom onset increased BDNF immunoreactivity in the affected area: the ventral horn of the spinal cord in neurons and notably also in non-neuronal cells [26]. Confirmation that the upregulation of BDNF was mediated through sigma-1 receptors via pharmacological antagonism or genetic manipulation (knockdown or knockout) has yet to be conducted in the ALS model and represents an important future study.

In vitro findings further support a role for sigma-1 receptor activity in regulating BDNF. Heat-induced aggregation of BDNF and GDNF were blocked by purified sigma-1 polypeptides [116]. In addition, the sigma-1 agonist SA4503 stimulated mature BDNF secretion from SH-SY5Y (human neuroblastoma) and B104 (rat neuroblastoma) cells, an effect that could be prevented with the sigma-1 antagonist NE100 [114]. Knockdown of sigma-1 receptors in B104 cells also decreased the ability of the cells to secrete mature BDNF, further underscoring a potential role for sigma-1 receptors in regulating BDNF processing and release [116].

Although not studied in neurodegeneration models, sigma-1 agonists have also been reported to stimulate nerve growth factor (NGF)-induced neurite outgrowth in cultured cells [117121] and potentiate epidermal growth factor (EGF)-induced neuritogenesis in PC12 (rat pheochromocytoma) cells overexpressing sigma-1 receptors [122]. In addition, sigma-1 activation with PRE084 enhanced neurite outgrowth in cerebellar granule cells via tropomyosin receptor kinase B (TrkB) signaling [123]. These observations raise the possibility that sigma-1 receptor activation has the potential to stimulate the activity of an array of neurotrophic factors to assist in recovery from CNS injury and disease.

10.4.2 Alterations in Neuronal Morphology

In many neurodegenerative conditions, aberrant neuronal morphology is observed. Dendritic and axonal deficiencies, in particular, are expected to compromise the integrity of neuronal connectivity within the CNS [124]. It is therefore noteworthy that sigma-1 receptors are found in key locations within neurons such as the growth cones [125], and agonists can promote neurite outgrowth from these sites through interactions with neurotrophic signaling pathways, as mentioned above. Additional studies have reported that decreased sigma-1 receptor expression can adversely affect dendritic arborization and axonal elongation in in vitro systems.

In hippocampal neurons, knockdown of sigma-1 receptors decreased dendritic arborization, diminished the formation and maturation of dendritic spines, and reduced protein markers of functional synapses [126]. Active forms of GTP (guanosine triphosphate)-binding Rac1 (ras-related C3 botulinum toxin substrate 1) and intact TIAM1 (T-cell lymphoma invasion and metastasis-inducing protein 1) in raft fractions were also reduced in sigma-1 receptor knockdown neurons [126], suggesting the contribution of this signaling pathway to the decreased dendritic arborization. The ability of a constitutively active Rac or caspase-3-resistant TIAM1 construct to rescue spine formation in sigma-1 knockdown neurons was supportive of such a role [126]. In addition, a free radical scavenger (N-acetylcysteine), superoxide dismutase activator (Tempol), or NOS inhibitor (nitro-L-arginine) was able to restore spine formation in sigma-1 receptor knockdown neurons [126]. Together, the data suggest that deficits in sigma-1 receptors can compromise dendritic spine formation and arborization through a free radical-sensitive mechanism involving the Rac1-GTP pathway [126].

Another study recently demonstrated that sigma-1 receptor depletion or ablation can also compromise axonal morphology. Sigma-1 receptor knockout mice exhibited lower densities of axons, as measured by actin neurofilament immunostaining in the cortex, when compared to wildtype mice [127]. In addition, a slower rate of degradation of p35 was observed when sigma-1 receptors were depleted by knockdown in vitro or knockout in vivo [127]. P35 is a major activator of cyclin-dependent kinase 5 (cdk5), which plays an important role in cytoskeletal dynamics of microtubules and actin neurofilaments [128]. In contrast, overexpression of sigma-1 receptors in CHO cells resulted in a faster rate of degradation of p35 [127]. The influence of sigma-1 receptors on p35 appears to involve indirect interactions since direct physical interactions were not detected [127]. Notably, myristate was shown to bind to sigma-1 receptors as a putative agonist, resulting in increased phosphorylation of actin neurofilament proteins and myristoylation of p35 within 24 h. This modification of p35 increases its susceptibility to protein degradation [129], thus ultimately eliciting axonal extension in wildtype neurons and rescuing deficits in axonal elongation in sigma-1 knockdown neurons [127]. Together, the data suggest that sigma-1 receptors, in response to stimulation by myristic acid, can influence axonal elongation by modulating p35 turnover.

Collectively, the data suggest that therapeutic interventions that restore sigma-1 receptor expression or stimulate its function can reverse alterations in neuronal structure or morphology that result from or are associated with disease. In this regard, it is noteworthy that administration of the sigma-1 agonist PRE084 restored the deficits in sigma-1 receptor expression in a cellular model of HD, which were accompanied by reductions in a multitude of neurodegenerative markers (Hyrskyluoto et al. 2013). Future investigations to characterize the relationship between sigma-1 receptors and morphological changes in the context of neurodegenerative disease models and treatment, particularly under in vivo conditions, would be of value.

10.4.3 Recruitment of New Cells to Damaged Area

Damage to the nervous system is characterized by the loss of neurons, as well as the recruitment of glial cells to the site of injury. In response to CNS injury, sigma-1 agonists have been reported to enhance microglial and astrocytic activities that are associated with repair (see Sects. 10.3.5 and 10.3.6, respectively). Although not yet studied in the context of neurodegeneration, there is some evidence from depression models that sigma-1 agonists promote neurogenesis. For example, the selective sigma-1 agonist SA4503 promoted neurogenesis following subchronic treatment in stress-naive rats [130] and also in calcium/calmodulin-dependent protein kinase IV null mice exhibiting depressive behaviors and impaired neurogenesis [131]. In another animal model of depression involving olfactory bulbectomized mice, sigma-1 active compounds such as dehydroepiandrosterone (DHEA) similarly enhanced neurogenesis [132], an effect that was inhibited by treatment with the sigma-1 antagonist NE-100. In contrast, knockout of sigma-1 receptors in mice suppressed neurite growth and the survival of newborn neuronal cells in the hippocampal dentate gyrus of adult mice [133]. Finally, stem cells which are under investigation for transplantation procedures [134] are enriched in sigma receptors [135].

10.5 Sigma-1 Receptor Activity as a Signal Amplifier in Neurodegeneration and Neuroprotection

Sigma ligands confer protective effects against many pathological mechanisms of neurodegeneration in preclinical studies, and have yielded great success as neuroprotective agents in different animal models of neurodegenerative disorders. However, the Food and Drug Administration (FDA) has yet to approve a selective sigma ligand for use in humans. In the single clinical study that tested a selective sigma ligand for the treatment of neurodegeneration, the selective sigma-1 agonist SA4503 failed to elicit significant functional recovery in the treated population compared to the placebo control after ischemic stroke [136]. However, post-hoc analysis of moderately and severely affected subjects showed significantly greater National Institutes of Health Stroke Scale improvements in the SA5403-treated group when compared with placebo (P=0.034 and P=0.038, respectively) [136]. Further clinical trials will be needed to optimize patient characteristics to identify a potential responder population, determine appropriate timing for treatment initiation and treatment duration, and evaluate the potential interaction of sigma-1 receptor therapy with other existing conventional pharmacological or non-pharmacological therapies.

It is noteworthy that many currently marketed psychotropic medications have significant affinity for sigma-1 receptors. Whether the therapeutic effects of these medications are mediated by sigma-1 receptor activity in humans remains unclear. Preclinical studies have shown that compounds such as fluvoxamine, DHEA-sulfate (DHEAS) and donepezil do elicit neuroprotective effects in part through activation of sigma-1 receptors, as their effects were attenuated with selective sigma-1 antagonists [13, 28, 137]. Below, we offer the hypothesis that alterations in the function or expression of sigma-1 receptors by themselves will likely be insufficient to cause disease or mitigate neurodegeneration. Rather, sigma-1 receptor dysfunction is likely to worsen disease progression, whereas stimulation may amplify pre-existing functional mechanisms of neuroprotection and/or restoration to slow down the disease progression.

10.5.1 Aberrant Sigma-1 Expression/Structure and Pathogenesis of Neurodegenerative Disease

Several recent studies have shown that decreased expression of sigma-1 receptors may contribute to the pathogenesis of neurodegenerative diseases. Mishina and colleagues, for example, used positron emission tomography (PET) with [11C] SA4530 to demonstrate reduced densities of sigma-1 receptors in the frontal, temporal and occipital lobes, cerebellum, and thalamus of early AD patients [138]. A later study in PD patients showed that the binding potential of the PET ligand [11C]SA4503 to sigma-1 receptors was significantly lower on the more affected than the less affected side of the anterior putamen [139]. However, there was no significant difference with respect to the binding potential between the patients and controls [139]. This supports the model that dysfunction in sigma-1 receptor protein expression augments, rather than initiates, disease progression.

Additionally, mutations in the sigma-1 receptor gene have been reported in ALS [93, 94], and sigma-1 receptor accumulates in intracellular protein aggregates in various neurodegenerative disorders, including trans-activation response DNA protein 43 proteinopathy, tauopathy, alpha-synucleinopathy, polyglutamine disease and intranuclear inclusion body disease [82, 140]. Since sigma-1 receptors have chaperone and regulatory roles [76, 84, 141], this accumulation may reflect a failed adaptive response to clear the inclusions during the course of the various diseases. It remains to be determined, however, whether these accumulations of sigma-1 receptors represent accumulations of unfolded/non-functional or functional sigma-1 receptors. This accumulation may also contribute to disease progression by limiting the number of soluble sigma-1 receptors present in the cell, which may in turn potentiate ER stress and subsequent apoptosis. This suggests that sigma-1 receptor dysfunction is a later effect in the pathologic process, after neurodegeneration has begun but possibly before the manifestation in clinical symptoms. Targeting the remaining functional sigma-1 receptors with sigma-1 agonists may therefore slow down disease progression (see Sect. 10.5.2).

Evidence from sigma-1 receptor knockdown and knockout studies in cellular and animal studies further support this hypothesized contributory role of sigma-1 receptors. In CHO cells, the knockdown of sigma-1 receptors destabilized the conformation of IRE1 and decreased cell survival following administration of the ER stressor thapsigargin [84]. In contrast, under vehicle conditions (in the absence of thapsigargin), sigma-1 knockdown had no effect on the stability of IRE1 or apoptosis [84]. In animal studies, Langa and colleagues demonstrated that homozygous mutant mice (mouse sigma-1 receptor gene, mSR1−/−) were viable and fertile with negligible overt phenotype compared with their wildtype littermates [142]. Mavlyutov et al. found that knockout of sigma-1 receptors produced slight motor abnormalities on a rotarod test, but did not itself result in an ALS phenotype or increased weight loss [143]. On the other hand, knockout of sigma-1 receptors in a SOD1-G93A mouse model of ALS exacerbated weight loss, produced an early decline in swimming performance, and ultimately decreased longevity [25].

Supporting the hypothesis that the deleterious consequences of sigma-1 receptor malfunction or aberrant expression manifest primarily under conditions of stress, retinal development also appears normal in sigma-1 knockout mice, with significant deficits (e.g., increased RGC loss and increased intraocular pressure) observed only with advanced age [144, 145]. A recent study showed that RGC death is accelerated in sigma-1 receptor knockout mice compared to wildtype following optic nerve crush, a model system for triggering apoptotic responses similar to those seen in glaucoma [144]. More extensive characterization has also been performed in sigma-1 receptor knockout mice with streptozotocin (STZ)-induced diabetes. Similar to the ocular crush model, STZ treatment accelerated retinal damage in sigma-1 receptor knockout mice; diabetic sigma-1 knockout mice showed fewer RGCs and more caspase-3 positive cells compared to non-diabetic wildtype mice, while sigma-1 knockout alone had no effects [145]. Additionally, relative to the other groups tested (non-diabetic knockout, non-diabetic wildtype and diabetic wildtype), diabetic sigma-1 receptor knockout mice showed increased intraocular pressure and deficits in scotopic threshold responses, which are the most sensitive electroretinogram responses observable with dim stimuli in the dark-adapted state and reflect RGC health [145].

10.5.2 Sigma-1 Activation as Intracellular Amplifier of Pre-existing Neuroprotective Mechanisms

Su and Hayashi proposed that sigma-1 receptors act as intracellular amplifiers for signal transduction, describing the biochemical actions of sigma-1 receptors as modulatory in nature and that the functional implications of these receptors may only be manifested when another biological system is first activated [146]. Consistent with this, the selective sigma-1 agonist (+)-pentazocine prolonged the association of sigma-1 receptors with IP3 receptors under ER stress but had no effect under normal conditions [76]. Likewise, sigma-1 agonists, without effects by themselves, potentiated bradykinin-induced alterations in cytosolic free Ca+2 concentrations [147]. In addition, Monnet and colleagues showed that in anesthetized rats, (+)-pentazocine had no detectable effects on its own, but potentiated NMDA-mediated glutamatergic stimulation [148, 149].

Relevant to neurodegenerative diseases, abnormal intracellular sigma-1 protein aggregates have been reported in various disorders, as mentioned in Sect. 10.5.1. Specific to ALS, sigma-1 receptor accumulation has been observed in lumbar alpha motor neurons of ALS patients and SOD1-G93A mice, cultured fibroblasts from ALS-8 patients with the P56S-VABP (vesicle-associated membrane protein-associated protein B) mutation, and in NSC34 (mouse motor neuron-like hybrid) cells transfected with the P56S-VABP mutation [140]. These accumulations co-localized with VAPB in the fibroblasts and NSC34 cells with the P56S-VABP mutation [140]. VABP is another ER protein, in which the P56S point mutation causes severe misfolding of the peptide and leads to the formation of cytoplasmic inclusion bodies and familial ALS [140]. Importantly, activation of sigma-1 receptors by PRE084 in P56S-VABP NSC34 cells ameliorated mutant VAPB aggregation and increased the degradation of soluble mutant VAPB without affecting the normal level of the wildtype proteins [140]. These results suggest targeting sigma-1 receptors with agonists can help ameliorate protein aggregation and inhibit disease progression by enhancing their innate chaperone activity.

Relevant to the neurorestorative potential of sigma-1 receptors, in PC12 cells, several sigma-1 agonists including (+)-pentazocine, imipramine, fluvoxamine and donepezil showed no effects on their own but potentiated NGF-induced neurite outgrowth [117120]. Co-administration of the sigma-1 antagonist NE-100 blocked this effect, confirming the specificity of sigma-1 receptor involvement [117120]. Moreover, the overexpression of sigma-1 receptors enhanced the NGF-induced neurite sprouting, while antisense deoxyoligonucleotides directed against sigma-1 receptors attenuated the NGF-induced neurite outgrowth [120].

10.5.3 Proposed Use of Sigma-1 Ligands in a Multi-target Therapeutic Approach

Due to the intrinsic modulatory role of sigma-1 receptors in disease and therapy, sigma-1 receptor activation as a stand-alone treatment appears unlikely to be sufficient to elicit observable clinical outcomes. The large body of preclinical evidence using primarily selective sigma-1 compounds described above indicates sigma-1 receptors are viable targets for therapeutic applications for CNS-related disorders. However, to combat the complex and multi-dimensional nature of neurodegenerative diseases, a multi-treatment approach would likely be most beneficial. Sigma-1 ligands, with the ability to affect multiple mechanisms and neural cell types that contribute to neurodegeneration through sigma-1 receptor activation, may therefore offer greater promise as an adjunct therapy. As we mentioned above, though there are no currently approved selective sigma-1 compounds for use in humans, many currently available psychotropic drugs interact with sigma-1 receptors. Using PET with [11C]SA4503, Ishikawa and colleagues showed that fluvoxamine, which has the highest affinity for sigma-1 receptors among SSRIs, binds to sigma-1 receptors in living human brains at therapeutic doses [150]. A follow-up study also showed that the acetylcholinesterase inhibitor donepezil binds to sigma-1 receptors at therapeutic doses [151]. Recently, in a mouse model of amyloid beta (25–35)-induced memory impairments, Maurice showed that protection by the sigma-1 agonist PRE084 is synergistic with donepezil [152]. Therefore, the repurposing or development of sigma-1 receptor active drugs, selective or not, requires further investigation as viable therapeutic approaches for treating neurodegenerative diseases. Moreover, the usage of selective sigma-1 ligands as an adjunct (vs. standalone) treatment may prove more fruitful in clinical trials and serve to validate the potential therapeutic significance of sigma-1 receptors as amplifiers of neuroprotective actions.

10.6 Conclusion

Sigma-1 receptors, with their wide range of effects on multiple signaling pathways, appear to be promising, druggable targets to help combat the complex pathophysiology of neurodegenerative disorders. In its apparent role as an intracellular amplifier, however, sigma-1 receptor activation will likely be most effective in a multi-target therapeutic approach in conjunction with other pharmacological interventions. Further understanding the signaling cascades regulated by sigma-1 receptors will aid in the development of novel therapies to slow the progression of neurodegeneration and/or reverse existing pathologies.

Contributor Information

Linda Nguyen, Department of Behavioral Medicine and Psychiatry, School of Medicine, West Virginia University, 930 Chestnut Ridge Road, Morgantown, WV 26506, USA; Department of Pharmaceutical Sciences, School of Pharmacy, One Medical Center, West Virginia University, Morgantown, WV 26506, USA.

Brandon P. Lucke-Wolds, Graduate Program in Neuroscience, School of Medicine, West Virginia University, One Medical Center Drive, Morgantown, WV 26506, USA

Shona Mookerjee, College of Pharmacy, Touro University California, 1310 Club Drive, Vallejo, CA 94592, USA.

Nidhi Kaushal, Takeda California, Inc., San Diego, CA 92121, USA.

Rae R. Matsumoto, Department of Behavioral Medicine and Psychiatry, School of Medicine, West Virginia University, 930 Chestnut Ridge Road, Morgantown, WV 26506, USA College of Pharmacy, Touro University California, 1310 Club Drive, Vallejo, CA 94592, USA.

References

  • 1.Prince M, et al. The Global Impact of Dementia. Alzheimer’s Disease International; 2015. World Alzheimer Report 2015. http://www.alz.co.uk/research/WorldAlzheimerReport2015.pdf. [Google Scholar]
  • 2.Villard V, et al. Antiamnesic and neuroprotective effects of the aminotetrahydrofuran derivative ANAVEX1-41 against amyloid beta(25–35)-induced toxicity in mice. Neuropsychopharmacology. 2009;34(6):1552–1566. doi: 10.1038/npp.2008.212. [DOI] [PubMed] [Google Scholar]
  • 3.Nguyen L, et al. Role of sigma-1 receptors in neurodegenerative diseases. J Pharmacol Sci. 2015;127(1):17–29. doi: 10.1016/j.jphs.2014.12.005. [DOI] [PubMed] [Google Scholar]
  • 4.Maurice T, Su TP. The pharmacology of sigma-1 receptors. Pharmacol Ther. 2009;124(2):195–206. doi: 10.1016/j.pharmthera.2009.07.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Su TP, et al. The sigma-1 receptor chaperone as an inter-organelle signaling modulator. Trends Pharmacol Sci. 2010;31(12):557–566. doi: 10.1016/j.tips.2010.08.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Urani A, et al. Enhanced antidepressant effect of sigma(1) (sigma(1)) receptor agonists in beta(25–35)-amyloid peptide-treated mice. Behav Brain Res. 2002;134(1–2):239–247. doi: 10.1016/s0166-4328(02)00033-5. [DOI] [PubMed] [Google Scholar]
  • 7.Matsumoto RR, Bowen WD, Su TP, editors. Sigma receptors: chemistry, cell biology and clinical implications. Springer; New York: 2007. p. 399. [Google Scholar]
  • 8.Aarts MM, Arundine M, Tymianski M. Novel concepts in excitotoxic neurodegeneration after stroke. Expert Rev Mol Med. 2003;5(30):1–22. doi: 10.1017/S1462399403007087. [DOI] [PubMed] [Google Scholar]
  • 9.Lai TW, Zhang S, Wang YT. Excitotoxicity and stroke: identifying novel targets for neuroprotection. Prog Neurobiol. 2014;115:157–188. doi: 10.1016/j.pneurobio.2013.11.006. [DOI] [PubMed] [Google Scholar]
  • 10.Vagnerova K, et al. Sigma 1 receptor agonists act as neuroprotective drugs through inhibition of inducible nitric oxide synthase. Anesth Analg. 2006;103(2):430–434. doi: 10.1213/01.ane.0000226133.85114.91. table of contents. [DOI] [PubMed] [Google Scholar]
  • 11.Ruscher K, et al. Effects of the sigma-1 receptor agonist 1-(3,4-dimethoxyphenethyl)-4-(3-phenylpropyl)-piperazine dihydro-chloride on inflammation after stroke. PLoS One. 2012;7(9):e45118. doi: 10.1371/journal.pone.0045118. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Allahtavakoli M, Jarrott B. Sigma-1 receptor ligand PRE-084 reduced infarct volume, neurological deficits, pro-inflammatory cytokines and enhanced anti-inflammatory cytokines after embolic stroke in rats. Brain Res Bull. 2011;85(3–4):219–224. doi: 10.1016/j.brainresbull.2011.03.019. [DOI] [PubMed] [Google Scholar]
  • 13.Sato S, et al. Antidepressant fluvoxamine reduces cerebral infarct volume and ameliorates sensorimotor dysfunction in experimental stroke. Neuroreport. 2014;25(10):731–736. doi: 10.1097/WNR.0000000000000162. [DOI] [PubMed] [Google Scholar]
  • 14.Ajmo CT, Jr, et al. Sigma receptor activation reduces infarct size at 24 hours after permanent middle cerebral artery occlusion in rats. Curr Neurovasc Res. 2006;3(2):89–98. doi: 10.2174/156720206776875849. [DOI] [PubMed] [Google Scholar]
  • 15.Katnik C, et al. Treatment with afobazole at delayed time points following ischemic stroke improves long-term functional and histological outcomes. Neurobiol Dis. 2014;62:354–364. doi: 10.1016/j.nbd.2013.10.011. [DOI] [PubMed] [Google Scholar]
  • 16.Takahashi H, et al. PPBP [4-phenyl-1-(4-phenylbutyl) piperidine] decreases brain injury after transient focal ischemia in rats. Stroke. 1996;27(11):2120–2123. doi: 10.1161/01.str.27.11.2120. [DOI] [PubMed] [Google Scholar]
  • 17.Lysko PG, et al. Neuroprotective effects of SKF 10,047 in cultured rat cerebellar neurons and in gerbil global brain ischemia. Stroke. 1992;23(3):414–419. doi: 10.1161/01.str.23.3.414. [DOI] [PubMed] [Google Scholar]
  • 18.Takahashi H, et al. PPBP [4-phenyl-1-(4-phenylbutyl) piperidine], a potent sigma-receptor ligand, decreases brain injury after transient focal ischemia in cats. Stroke. 1995;26(9):1676–1682. doi: 10.1161/01.str.26.9.1676. [DOI] [PubMed] [Google Scholar]
  • 19.Ruscher K, et al. The sigma-1 receptor enhances brain plasticity and functional recovery after experimental stroke. Brain. 2011;134(Pt 3):732–746. doi: 10.1093/brain/awq367. [DOI] [PubMed] [Google Scholar]
  • 20.Hazell AS. Excitotoxic mechanisms in stroke: an update of concepts and treatment strategies. Neurochem Int. 2007;50(7–8):941–953. doi: 10.1016/j.neuint.2007.04.026. [DOI] [PubMed] [Google Scholar]
  • 21.Jarrott B, Williams SJ. Chronic brain inflammation: the neurochemical basis for drugs to reduce inflammation. Neurochem Res. 2015;41:523. doi: 10.1007/s11064-015-1661-7. [DOI] [PubMed] [Google Scholar]
  • 22.Yang ZJ, et al. Sigma receptor ligand 4-pheny l-1-(4-phenylbutyl)-piperidine modulates neuronal nitric oxide synthase/postsynaptic density-95 coupling mechanisms and protects against neonatal ischemic degeneration of striatal neurons. Exp Neurol. 2010;221(1):166–174. doi: 10.1016/j.expneurol.2009.10.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Griesmaier E, et al. Neuroprotective effects of the sigma-1 receptor ligand PRE-084 against excitotoxic perinatal brain injury in newborn mice. Exp Neurol. 2012;237(2):388–395. doi: 10.1016/j.expneurol.2012.06.030. [DOI] [PubMed] [Google Scholar]
  • 24.Mancuso R, et al. Sigma-1R agonist improves motor function and motoneuron survival in ALS mice. Neurotherapeutics. 2012;9(4):814–826. doi: 10.1007/s13311-012-0140-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Mavlyutov TA, et al. Lack of sigma-1 receptor exacerbates ALS progression in mice. Neuroscience. 2013;240:129–134. doi: 10.1016/j.neuroscience.2013.02.035. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Peviani M, et al. Neuroprotective effects of the sigma-1 receptor (S1R) agonist PRE-084, in a mouse model of motor neuron disease not linked to SOD1 mutation. Neurobiol Dis. 2014;62:218–232. doi: 10.1016/j.nbd.2013.10.010. [DOI] [PubMed] [Google Scholar]
  • 27.Francardo V, et al. Pharmacological stimulation of sigma-1 receptors has neurorestorative effects in experimental parkinsonism. Brain. 2014;137:1998. doi: 10.1093/brain/awu107. [DOI] [PubMed] [Google Scholar]
  • 28.Meunier J, Ieni J, Maurice T. The anti-amnesic and neuroprotective effects of donepezil against amyloid beta25–35 peptide-induced toxicity in mice involve an interaction with the sigma1 receptor. Br J Pharmacol. 2006;149(8):998–1012. doi: 10.1038/sj.bjp.0706927. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Maurice T, Su TP, Privat A. Sigma1 (sigma 1) receptor agonists and neurosteroids attenuate B25–35-amyloid peptide-induced amnesia in mice through a common mechanism. Neuroscience. 1998;83(2):413–428. doi: 10.1016/s0306-4522(97)00405-3. [DOI] [PubMed] [Google Scholar]
  • 30.Schetz JA, et al. A prototypical sigma-1 receptor antagonist protects against brain ischemia. Brain Res. 2007;1181:1–9. doi: 10.1016/j.brainres.2007.08.068. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Matsumoto RR, et al. Sigma (sigma) receptors as potential therapeutic targets to mitigate psychostimulant effects. Adv Pharmacol. 2014;69:323–386. doi: 10.1016/B978-0-12-420118-7.00009-3. [DOI] [PubMed] [Google Scholar]
  • 32.Diaz JL, et al. Selective sigma-1 (sig1) receptor antagonists: emerging target for the treatment of neuropathic pain. Cent Nerv Syst Agents Med Chem (Formerly Current Medicinal Chemistry-Central Nervous System Agents) 2009;9(3):172–183. doi: 10.2174/1871524910909030172. [DOI] [PubMed] [Google Scholar]
  • 33.Hong J, et al. Sigma-1 receptor deficiency reduces MPTP-induced parkinsonism and death of dopaminergic neurons. Cell Death Dis. 2015;6(7):e1832. doi: 10.1038/cddis.2015.194. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Sheldon AL, Robinson MB. The role of glutamate transporters in neurodegenerative diseases and potential opportunities for intervention. Neurochem Int. 2007;51(6–7):333–355. doi: 10.1016/j.neuint.2007.03.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Lau A, Tymianski M. Glutamate receptors, neurotoxicity and neurodegeneration. Pflugers Arch. 2010;460(2):525–542. doi: 10.1007/s00424-010-0809-1. [DOI] [PubMed] [Google Scholar]
  • 36.Zündorf G, Reiser G. Calcium dysregulation and homeostasis of neural calcium in the molecular mechanisms of neurodegenerative diseases provide multiple targets for neuroprotection. Antioxid Redox Signal. 2011;14(7):1275–1288. doi: 10.1089/ars.2010.3359. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Mattson MP. Calcium and neurodegeneration. Aging Cell. 2007;6(3):337–350. doi: 10.1111/j.1474-9726.2007.00275.x. [DOI] [PubMed] [Google Scholar]
  • 38.Hetz C, Mollereau B. Disturbance of endoplasmic reticulum proteostasis in neurodegenerative diseases. Nat Rev Neurosci. 2014;15(4):233–249. doi: 10.1038/nrn3689. [DOI] [PubMed] [Google Scholar]
  • 39.Sovolyova N, et al. Stressed to death - mechanisms of ER stress-induced cell death. Biol Chem. 2014;395(1):1–13. doi: 10.1515/hsz-2013-0174. [DOI] [PubMed] [Google Scholar]
  • 40.Halliday M, Mallucci GR. Targeting the unfolded protein response in neurodegeneration: a new approach to therapy. Neuropharmacology. 2014;76(Pt A):169–174. doi: 10.1016/j.neuropharm.2013.08.034. [DOI] [PubMed] [Google Scholar]
  • 41.Chaturvedi RK, Flint Beal M. Mitochondrial diseases of the brain. Free Radic Biol Med. 2013;63:1–29. doi: 10.1016/j.freeradbiomed.2013.03.018. [DOI] [PubMed] [Google Scholar]
  • 42.Perry VH, Nicoll JA, Holmes C. Microglia in neurodegenerative disease. Nat Rev Neurol. 2010;6(4):193–201. doi: 10.1038/nrneurol.2010.17. [DOI] [PubMed] [Google Scholar]
  • 43.Frank-Cannon TC, et al. Does neuroinflammation fan the flame in neurodegenerative diseases. Mol Neurodegener. 2009;4(47):1–13. doi: 10.1186/1750-1326-4-47. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Maragakis NJ, Rothstein JD. Mechanisms of disease: astrocytes in neurodegenerative disease. Nat Clin Pract Neurol. 2006;2(12):679–689. doi: 10.1038/ncpneuro0355. [DOI] [PubMed] [Google Scholar]
  • 45.Lu B, et al. BDNF-based synaptic repair as a disease-modifying strategy for neurodegenerative diseases. Nat Rev Neurosci. 2013;14(6):401–416. doi: 10.1038/nrn3505. [DOI] [PubMed] [Google Scholar]
  • 46.Weissmiller AM, Wu C. Current advances in using neurotrophic factors to treat neurodegenerative disorders. Transl Neurodegeneration. 2012;1:14. doi: 10.1186/2047-9158-1-14. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Siegel GJ, Chauhan NB. Neurotrophic factors in Alzheimer’s and Parkinson’s disease brain. Brain Res Rev. 2000;33(2–3):199–227. doi: 10.1016/s0165-0173(00)00030-8. [DOI] [PubMed] [Google Scholar]
  • 48.Horner PJ, Gage FH. Regenerating the damaged central nervous system. Nature. 2000;407(6807):963–970. doi: 10.1038/35039559. [DOI] [PubMed] [Google Scholar]
  • 49.Szydlowska K, Tymianski M. Calcium, ischemia and excitotoxicity. Cell Calcium. 2010;47(2):122–129. doi: 10.1016/j.ceca.2010.01.003. [DOI] [PubMed] [Google Scholar]
  • 50.Hynd MR, Scott HL, Dodd PR. Glutamate-mediated excitotoxicity and neurodegeneration in Alzheimer’s disease. Neurochem Int. 2004;45(5):583–595. doi: 10.1016/j.neuint.2004.03.007. [DOI] [PubMed] [Google Scholar]
  • 51.Krasnova IN, Cadet JL. Methamphetamine toxicity and messengers of death. Brain Res Rev. 2009;60(2):379–407. doi: 10.1016/j.brainresrev.2009.03.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Farber NB, et al. Receptor mechanisms and circuitry underlying NMDA antagonist neurotoxicity. Mol Psychiatry. 2002;7(1):32–43. doi: 10.1038/sj.mp.4000912. [DOI] [PubMed] [Google Scholar]
  • 53.Martin PM, et al. The sigma receptor ligand (+)-pentazocine prevents apoptotic retinal ganglion cell death induced in vitro by homocysteine and glutamate. Brain Res Mol Brain Res. 2004;123(1–2):66–75. doi: 10.1016/j.molbrainres.2003.12.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.DeCoster MA, et al. Sigma receptor-mediated neuroprotection against glutamate toxicity in primary rat neuronal cultures. Brain Res. 1995;671(1):45–53. doi: 10.1016/0006-8993(94)01294-r. [DOI] [PubMed] [Google Scholar]
  • 55.Dun Y, et al. Prevention of excitotoxicity in primary retinal ganglion cells by (+)-pentazocine, a sigma receptor-1 specific ligand. Invest Ophthalmol Vis Sci. 2007;48(10):4785–4794. doi: 10.1167/iovs.07-0343. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Shen YC, et al. Dimemorfan protects rats against ischemic stroke through activation of sigma-1 receptor-mediated mechanisms by decreasing glutamate accumulation. J Neurochem. 2008;104(2):558–572. doi: 10.1111/j.1471-4159.2007.05058.x. [DOI] [PubMed] [Google Scholar]
  • 57.Shimazu S, et al. Sigma receptor ligands attenuate N-methyl-D-aspartate cytotoxicity in dopaminergic neurons of mesencephalic slice cultures. Eur J Pharmacol. 2000;388(2):139–146. doi: 10.1016/s0014-2999(99)00852-3. [DOI] [PubMed] [Google Scholar]
  • 58.Fu Y, et al. Fluvoxamine increased glutamate release by activating both 5-HT(3) and sigma-1 receptors in prelimbic cortex of chronic restraint stress C57BL/6 mice. Biochim Biophys Acta. 2012;1823(4):826–837. doi: 10.1016/j.bbamcr.2012.01.008. [DOI] [PubMed] [Google Scholar]
  • 59.Lu CW, et al. Sigma-1 receptor agonist SKF10047 inhibits glutamate release in rat cerebral cortex nerve endings. J Pharmacol Exp Ther. 2012;341(2):532–542. doi: 10.1124/jpet.111.191189. [DOI] [PubMed] [Google Scholar]
  • 60.Balasuriya D, Stewart AP, Edwardson JM. The sigma-1 receptor interacts directly with GluN1 but not GluN2A in the GluN1/GluN2A NMDA receptor. J Neurosci. 2013;33(46):18219–18224. doi: 10.1523/JNEUROSCI.3360-13.2013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Pabba M, Sibille E. Sigma-1 and N-methyl-D-aspartate receptors: a partnership with beneficial outcomes. Mol Neuropsychiatry. 2015;1(1):47–51. doi: 10.1159/000376549. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Martina M, et al. The sigma-1 receptor modulates NMDA receptor synaptic transmission and plasticity via SK channels in rat hippocampus. J Physiol. 2007;578(Pt 1):143–157. doi: 10.1113/jphysiol.2006.116178. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Rodriguez-Munoz M, et al. The sigma1 receptor engages the redox-regulated HINT1 protein to bring opioid analgesia under NMDA receptor negative control. Antioxid Redox Signal. 2015;22(10):799–818. doi: 10.1089/ars.2014.5993. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Pabba M, et al. NMDA receptors are upregulated and trafficked to the plasma membrane after sigma-1 receptor activation in the rat hippocampus. J Neurosci. 2014;34(34):11325–11338. doi: 10.1523/JNEUROSCI.0458-14.2014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Roh DH, et al. Sigma-1 receptor-induced increase in murine spinal NR1 phosphorylation is mediated by the PKCalpha and epsilon, but not the PKCzeta, isoforms. Neurosci Lett. 2010;477(2):95–99. doi: 10.1016/j.neulet.2010.04.041. [DOI] [PubMed] [Google Scholar]
  • 66.Rodriguez-Munoz M, et al. The ON:OFF switch, sigma1R-HINT1 protein, controls GPCR-NMDA receptor cross-regulation: implications in neurological disorders. Oncotarget. 2015;6(34):35458–35477. doi: 10.18632/oncotarget.6064. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Xu Q, et al. Sigma 1 receptor activation regulates brain-derived neurotrophic factor through NR2A-CaMKIV-TORC1 pathway to rescue the impairment of learning and memory induced by brain ischaemia/reperfusion. Psychopharmacology. 2015;232(10):1779–1791. doi: 10.1007/s00213-014-3809-6. [DOI] [PubMed] [Google Scholar]
  • 68.Kim HC, et al. Carbetapentane attenuates kainate-induced seizures via sigma-1 receptor modulation. Life Sci. 2001;69(8):915–922. doi: 10.1016/s0024-3205(01)01181-x. [DOI] [PubMed] [Google Scholar]
  • 69.Shin EJ, et al. The dextromethorphan analog dimemorfan attenuates kainate-induced seizures via sigma1 receptor activation: comparison with the effects of dextromethorphan. Br J Pharmacol. 2005;144(7):908–918. doi: 10.1038/sj.bjp.0705998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Tuerxun T, et al. SA4503, a sigma-1 receptor agonist, prevents cultured cortical neurons from oxidative stress-induced cell death via suppression of MAPK pathway activation and glutamate receptor expression. Neurosci Lett. 2010;469(3):303–308. doi: 10.1016/j.neulet.2009.12.013. [DOI] [PubMed] [Google Scholar]
  • 71.Kumamaru E, et al. Glucocorticoid prevents brain-derived neurotrophic factor-mediated maturation of synaptic function in developing hippocampal neurons through reduction in the activity of mitogen-activated protein kinase. Mol Endocrinol. 2008;22(3):546–558. doi: 10.1210/me.2007-0264. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Tchedre KT, Yorio T. Sigma-1 receptors protect RGC-5 cells from apoptosis by regulating intracellular calcium, Bax levels, and caspase-3 activation. Invest Ophthalmol Vis Sci. 2008;49(6):2577–2588. doi: 10.1167/iovs.07-1101. [DOI] [PubMed] [Google Scholar]
  • 73.Behensky AA, et al. Afobazole activation of sigma-1 receptors modulates neuronal responses to amyloid-beta25–35. J Pharmacol Exp Ther. 2013;347(2):468–477. doi: 10.1124/jpet.113.208330. [DOI] [PubMed] [Google Scholar]
  • 74.Tsai SY, et al. Sigma-1 receptor chaperones in neurodegenerative and psychiatric disorders. Expert Opin Ther Targets. 2014;18(12):1461–1476. doi: 10.1517/14728222.2014.972939. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Monnet FP. Sigma-1 receptor as regulator of neuronal intracellular Ca2+: clinical and therapeutic relevance. Biol Cell. 2005;97(12):873–883. doi: 10.1042/BC20040149. [DOI] [PubMed] [Google Scholar]
  • 76.Hayashi T, Su TP. Sigma-1 receptor chaperones at the ER-mitochondrion interface regulate Ca(2+) signaling and cell survival. Cell. 2007;131(3):596–610. doi: 10.1016/j.cell.2007.08.036. [DOI] [PubMed] [Google Scholar]
  • 77.Tsai SY, et al. Sigma-1 receptor chaperones and diseases. Cent Nerv Syst Agents Med Chem. 2009;9(3):184–189. doi: 10.2174/1871524910909030184. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Herrera Y, et al. Sigma-1 receptor modulation of acid-sensing ion channel a (ASIC1a) and ASIC1a-induced Ca2+ influx in rat cortical neurons. J Pharmacol Exp Ther. 2008;327(2):491–502. doi: 10.1124/jpet.108.143974. [DOI] [PubMed] [Google Scholar]
  • 79.Tchedre KT, et al. Sigma-1 receptor regulation of voltage-gated calcium channels involves a direct interaction. Invest Ophthalmol Vis Sci. 2008;49(11):4993–5002. doi: 10.1167/iovs.08-1867. [DOI] [PubMed] [Google Scholar]
  • 80.Hall AA, et al. Sigma receptors suppress multiple aspects of microglial activation. Glia. 2009;57(7):744–754. doi: 10.1002/glia.20802. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81.Walter P, Ron D. The unfolded protein response: from stress pathway to homeostatic regulation. Science. 2011;334(6059):1081–1086. doi: 10.1126/science.1209038. [DOI] [PubMed] [Google Scholar]
  • 82.Miki Y, et al. Accumulation of the sigma-1 receptor is common to neuronal nuclear inclusions in various neurodegenerative diseases. Neuropathology. 2014;34(2):148–158. doi: 10.1111/neup.12080. [DOI] [PubMed] [Google Scholar]
  • 83.Ha Y, et al. Sigma receptor 1 modulates endoplasmic reticulum stress in retinal neurons. Invest Ophthalmol Vis Sci. 2011;52(1):527–540. doi: 10.1167/iovs.10-5731. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.Mori T, et al. Sigma-1 receptor chaperone at the ER-mitochondrion interface mediates the mitochondrion-ER-nucleus signaling for cellular survival. PLoS One. 2013;8(10):e76941. doi: 10.1371/journal.pone.0076941. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.Chu U, Ruoho AE. Biochemical pharmacology of the sigma-1 receptor. Mol Pharmacol. 2015;89:142. doi: 10.1124/mol.115.101170. [DOI] [PubMed] [Google Scholar]
  • 86.Ron D, Walter P. Signal integration in the endoplasmic reticulum unfolded protein response. Nat Rev Mol Cell Biol. 2007;8(7):519–529. doi: 10.1038/nrm2199. [DOI] [PubMed] [Google Scholar]
  • 87.Mitsuda T, et al. Sigma-1Rs are upregulated via PERK/eIF2α/ATF4 pathway and execute protective function in ER stress. Biochem Biophys Res Commun. 2011;415(3):519–525. doi: 10.1016/j.bbrc.2011.10.113. [DOI] [PubMed] [Google Scholar]
  • 88.Omi T, et al. Fluvoxamine alleviates ER stress via induction of sigma-1 receptor. Cell Death Dis. 2014;5:e1332. doi: 10.1038/cddis.2014.301. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Shimazawa M, et al. Effect of a sigma-1 receptor agonist, cutamesine dihydrochloride (SA4503), on photoreceptor cell death against light-induced damage. Exp Eye Res. 2015;132:64–72. doi: 10.1016/j.exer.2015.01.017. [DOI] [PubMed] [Google Scholar]
  • 90.Narita N, et al. Interactions of selective serotonin reuptake inhibitors with subtypes of σ receptors in rat brain. Eur J Pharmacol. 1996;307(1):117–119. doi: 10.1016/0014-2999(96)00254-3. [DOI] [PubMed] [Google Scholar]
  • 91.Wu Z, Bowen WD. Role of sigma-1 receptor C-terminal segment in inositol 1,4,5-trisphosphate receptor activation: constitutive enhancement of calcium signaling in MCF-7 tumor cells. J Biol Chem. 2008;283(42):28198–28215. doi: 10.1074/jbc.M802099200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Shioda N, et al. Expression of a truncated form of the endoplasmic reticulum chaperone protein, sigma1 receptor, promotes mitochondrial energy depletion and apoptosis. J Biol Chem. 2012;287(28):23318–23331. doi: 10.1074/jbc.M112.349142. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Al-Saif A, Al-Mohanna F, Bohlega S. A mutation in sigma-1 receptor causes juvenile amyotrophic lateral sclerosis. Ann Neurol. 2011;70(6):913–919. doi: 10.1002/ana.22534. [DOI] [PubMed] [Google Scholar]
  • 94.Luty AA, et al. Sigma nonopioid intracellular receptor 1 mutations cause frontotemporal lobar degeneration-motor neuron disease. Ann Neurol. 2010;68(5):639–649. doi: 10.1002/ana.22274. [DOI] [PubMed] [Google Scholar]
  • 95.Wegleiter K, et al. The sigma-1 receptor agonist 4-phenyl-1-(4-phenylbutyl)piperidine (PPBP) protects against newborn excitotoxic brain injury by stabilizing the mitochondrial membrane potential in vitro and inhibiting microglial activation in vivo. Exp Neurol. 2014;261:501–509. doi: 10.1016/j.expneurol.2014.07.022. [DOI] [PubMed] [Google Scholar]
  • 96.Klouz A, et al. Protection of cellular and mitochondrial functions against liver ischemia by N-benzyl-N′-(2-hydroxy-3,4-dimethoxybenzyl)-piperazine (BHDP), a sigma1 ligand. Eur J Pharmacol. 2008;578(2–3):292–299. doi: 10.1016/j.ejphar.2007.09.038. [DOI] [PubMed] [Google Scholar]
  • 97.Burke NN, et al. Minocycline modulates neuropathic pain behaviour and cortical M1-M2 microglial gene expression in a rat model of depression. Brain Behav Immun. 2014;42:147–156. doi: 10.1016/j.bbi.2014.06.015. [DOI] [PubMed] [Google Scholar]
  • 98.Robson MJ, et al. SN79, a sigma receptor ligand, blocks methamphetamine-induced microglial activation and cytokine upregulation. Exp Neurol. 2013;247:134–142. doi: 10.1016/j.expneurol.2013.04.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Wu Z, et al. Allosteric modulation of sigma-1 receptors by SKF83959 inhibits microglia-mediated inflammation. J Neurochem. 2015;134(5):904–914. doi: 10.1111/jnc.13182. [DOI] [PubMed] [Google Scholar]
  • 100.Cuevas J, et al. Afobazole modulates microglial function via activation of both sigma-1 and sigma-2 receptors. J Pharmacol Exp Ther. 2011;339(1):161–172. doi: 10.1124/jpet.111.182816. [DOI] [PubMed] [Google Scholar]
  • 101.Dong H, et al. Sigma-1 receptor modulates neuroinflammation after traumatic brain injury. Cell Mol Neurobiol. 2015;36:639. doi: 10.1007/s10571-015-0244-0. [DOI] [PubMed] [Google Scholar]
  • 102.Behensky AA, et al. Stimulation of sigma receptors with afobazole blocks activation of microglia and reduces toxicity caused by amyloid-beta25–35. J Pharmacol Exp Ther. 2013;347(2):458–467. doi: 10.1124/jpet.113.208348. [DOI] [PubMed] [Google Scholar]
  • 103.Derocq JM, et al. In vivo inhibition of endotoxin-induced pro-inflammatory cytokines production by the sigma ligand SR 31747. J Pharmacol Exp Ther. 1995;272(1):224–230. [PubMed] [Google Scholar]
  • 104.Casellas P, et al. Immunopharmacological profile of SR 31747: in vitro and in vivo studies on humoral and cellular responses. J Neuroimmunol. 1994;52(2):193–203. doi: 10.1016/0165-5728(94)90113-9. [DOI] [PubMed] [Google Scholar]
  • 105.Oxombre B, et al. High-affinity sigma1 protein agonist reduces clinical and pathological signs of experimental autoimmune encephalomyelitis. Br J Pharmacol. 2015;172(7):1769–1782. doi: 10.1111/bph.13037. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.Brown A, McFarlin D, Raine C. Chronologic neuropathology of relapsing experimental allergic encephalomyelitis in the mouse. Lab Investig; A Journal of Technical Methods and Pathology. 1982;46(2):171–185. [PubMed] [Google Scholar]
  • 107.Koyama Y. Signaling molecules regulating phenotypic conversions of astrocytes and glial scar formation in damaged nerve tissues. Neurochem Int. 2014;78:35–42. doi: 10.1016/j.neuint.2014.08.005. [DOI] [PubMed] [Google Scholar]
  • 108.Penas C, et al. Sigma receptor agonist 2-(4-morpholinethyl)-1-phenylcyclohexanecarboxylate (PRE084) increases GDNF and BiP expres sion and promotes neuroprotection after root avulsion injury. J Neurotrauma. 2011;28(5):831–840. doi: 10.1089/neu.2010.1674. [DOI] [PubMed] [Google Scholar]
  • 109.Robson MJ, et al. SN79, a sigma receptor antagonist, attenuates methamphetamine-induced astrogliosis through a blockade of OSMR/gp130 signaling and STAT3 phosphorylation. Exp Neurol. 2014;254:180–189. doi: 10.1016/j.expneurol.2014.01.020. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Zamanian JL, et al. Genomic analysis of reactive astrogliosis. J Neurosci. 2012;32(18):6391–6410. doi: 10.1523/JNEUROSCI.6221-11.2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.Bothwell M. NGF, BDNF, NT3, and NT4. In: Lewin GR, Carter BD, editors. Neurotrophic factors. Springer; Berlin/Heidelberg: 2014. pp. 3–15. [Google Scholar]
  • 112.Budni J, et al. The involvement of BDNF, NGF and GDNF in aging and Alzheimer’s disease. Aging Dis. 2015;6(5):331–341. doi: 10.14336/AD.2015.0825. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113.Allen SJ, et al. GDNF, NGF and BDNF as therapeutic options for neurodegeneration. Pharmacol Ther. 2013;138(2):155–175. doi: 10.1016/j.pharmthera.2013.01.004. [DOI] [PubMed] [Google Scholar]
  • 114.Li L, et al. Rescue of adult mouse motoneurons from injury-induced cell death by glial cell line-derived neurotrophic factor. Proc Natl Acad Sci U S A. 1995;92(21):9771–9775. doi: 10.1073/pnas.92.21.9771. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.Tomac A, et al. Retrograde axonal transport of glial cell line-derived neurotrophic factor in the adult nigrostriatal system suggests a trophic role in the adult. Proc Natl Acad Sci U S A. 1995;92(18):8274–8278. doi: 10.1073/pnas.92.18.8274. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.Fujimoto M, et al. Sigma-1 receptor chaperones regulate the secretion of brain-derived neurotrophic factor. Synapse. 2012;66(7):630–639. doi: 10.1002/syn.21549. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.Ishima T, Fujita Y, Hashimoto K. Interaction of new antidepressants with sigma-1 receptor chaperones and their potentiation of neurite outgrowth in PC12 cells. Eur J Pharmacol. 2014;727:167–173. doi: 10.1016/j.ejphar.2014.01.064. [DOI] [PubMed] [Google Scholar]
  • 118.Ishima T, Hashimoto K. Potentiation of nerve growth factor-induced neurite outgrowth in PC12 cells by ifenprodil: the role of sigma-1 and IP3 receptors. PLoS One. 2012;7(5):e37989. doi: 10.1371/journal.pone.0037989. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119.Ishima T, et al. Potentiation of nerve growth factor-induced neurite outgrowth in PC12 cells by donepezil: role of sigma-1 receptors and IP3 receptors. Prog Neuro-Psychopharmacol Biol Psychiatry. 2008;32(7):1656–1659. doi: 10.1016/j.pnpbp.2008.06.011. [DOI] [PubMed] [Google Scholar]
  • 120.Takebayashi M, Hayashi T, Su TP. Nerve growth factor-induced neurite sprouting in PC12 cells involves sigma-1 receptors: implications for antidepressants. J Pharmacol Exp Ther. 2002;303(3):1227–1237. doi: 10.1124/jpet.102.041970. [DOI] [PubMed] [Google Scholar]
  • 121.Nishimura T, et al. 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(7):e2558. doi: 10.1371/journal.pone.0002558. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Takebayashi M, Hayashi T, Su TP. σ-1 Receptors potentiate epidermal growth factor signaling towards neuritogenesis in PC12 cells: potential relation to lipid raft reconstitution. Synapse. 2004;53(2):90–103. doi: 10.1002/syn.20041. [DOI] [PubMed] [Google Scholar]
  • 123.Kimura Y, et al. Sigma-1 receptor enhances neurite elongation of cerebellar granule neurons via TrkB signaling. PLoS One. 2013;8(10):e75760. doi: 10.1371/journal.pone.0075760. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124.Van Battum EY, Brignani S, Pasterkamp RJ. Axon guidance proteins in neurological disorders. Lancet Neurol. 2015;14(5):532–546. doi: 10.1016/S1474-4422(14)70257-1. [DOI] [PubMed] [Google Scholar]
  • 125.Hayashi T, Su T-P. Regulating ankyrin dynamics: roles of sigma-1 receptors. Proc Natl Acad Sci. 2001;98(2):491–496. doi: 10.1073/pnas.98.2.491. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 126.Tsai S-Y, et al. Sigma-1 receptors regulate hippocampal dendritic spine formation via a free radical-sensitive mechanism involving Rac1· GTP pathway. Proc Natl Acad Sci. 2009;106(52):22468–22473. doi: 10.1073/pnas.0909089106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 127.Tsai S-YA, et al. Sigma-1 receptor regulates Tau phosphorylation and axon extension by shaping p35 turnover via myristic acid. Proc Natl Acad Sci. 2015;112(21):6742–6747. doi: 10.1073/pnas.1422001112. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 128.Hahn CM, et al. Role of cyclin-dependent kinase 5 and its activator P35 in local axon and growth cone stabilization. Neuroscience. 2005;134(2):449–465. doi: 10.1016/j.neuroscience.2005.04.020. [DOI] [PubMed] [Google Scholar]
  • 129.Minegishi S, et al. Membrane association facilitates degradation and cleavage of the cyclin-dependent kinase 5 activators p35 and p39. Biochemistry. 2010;49(26):5482–5493. doi: 10.1021/bi100631f. [DOI] [PubMed] [Google Scholar]
  • 130.Lucas G, et al. Further evidence for an antidepressant potential of the selective sigma1 agonist SA 4503: electrophysiological, morphological and behavioural studies. Int J Neuropsychopharmacol. 2008;11(4):485–495. doi: 10.1017/S1461145708008547. [DOI] [PubMed] [Google Scholar]
  • 131.Fukunaga K, Sakagami H, Moriguchi S. Sigma-1 receptor agonist and fluvoxamine rescue depressive behaviors in CaMKIV null mice. FASEB J. 2015;29(1 Supplement):931. [Google Scholar]
  • 132.Moriguchi S, et al. Stimulation of the sigma-1 receptor by DHEA enhances synaptic efficacy and neurogenesis in the hippocampal dentate gyrus of olfactory bulbectomized mice. PLoS One. 2013;8(4):e60863. doi: 10.1371/journal.pone.0060863. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 133.Sha S, et al. Sigma-1 receptor knockout impairs neurogenesis in dentate gyrus of adult hippocampus via down-regulation of NMDA receptors. CNS Neurosci Ther. 2013;19(9):705–713. doi: 10.1111/cns.12129. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 134.Kerkis I, et al. Neural and mesenchymal stem cells in animal models of Huntington’s disease: past experiences and future challenges. Stem Cell Res Ther. 2015;6(1):1–15. doi: 10.1186/s13287-015-0248-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 135.Nixon CC, et al. Cocaine exposure impairs multineage hematopoiesis of human hematopoietic progenitor cells mediated by the sigma-1 receptor. Sci Rep. 2015;5:8670. doi: 10.1038/srep08670. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136.Urfer R, et al. Phase II trial of the sigma-1 receptor agonist cutamesine (SA4503) for recovery enhancement after acute ischemic stroke. Stroke. 2014;45(11):3304–3310. doi: 10.1161/STROKEAHA.114.005835. [DOI] [PubMed] [Google Scholar]
  • 137.Bucolo C, Drago F. Effects of neurosteroids on ischemia-reperfusion injury in the rat retina: role of sigma1 recognition sites. Eur J Pharmacol. 2004;498(1–3):111–114. doi: 10.1016/j.ejphar.2004.06.067. [DOI] [PubMed] [Google Scholar]
  • 138.Mishina M, et al. Low density of sigma1 receptors in early Alzheimer’s disease. Ann Nucl Med. 2008;22(3):151–156. doi: 10.1007/s12149-007-0094-z. [DOI] [PubMed] [Google Scholar]
  • 139.Mishina M, et al. Function of sigma1 receptors in Parkinson’s disease. Acta Neurol Scand. 2005;112(2):103–107. doi: 10.1111/j.1600-0404.2005.00432.x. [DOI] [PubMed] [Google Scholar]
  • 140.Prause J, et al. Altered localization, abnormal modification and loss of function of sigma receptor-1 in amyotrophic lateral sclerosis. Hum Mol Genet. 2013;22(8):1581–1600. doi: 10.1093/hmg/ddt008. [DOI] [PubMed] [Google Scholar]
  • 141.Hayashi T, et al. The lifetime of UDP-galactose:ceramide galactosyltransferase is controlled by a distinct endoplasmic reticulum-associated degradation (ERAD) regulated by sigma-1 receptor chaperones. J Biol Chem. 2012;287(51):43156–43169. doi: 10.1074/jbc.M112.380444. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 142.Langa F, et al. Generation and phenotypic analysis of sigma receptor type I (sigma 1) knockout mice. Eur J Neurosci. 2003;18(8):2188–2196. doi: 10.1046/j.1460-9568.2003.02950.x. [DOI] [PubMed] [Google Scholar]
  • 143.Mavlyutov TA, et al. The sigma-1 receptor is enriched in postsynaptic sites of C-terminals in mouse motoneurons. An anatomical and behavioral study. Neuroscience. 2010;167(2):247–255. doi: 10.1016/j.neuroscience.2010.02.022. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 144.Mavlyutov TA, Nickells RW, Guo LW. Accelerated retinal ganglion cell death in mice deficient in the sigma-1 receptor. Mol Vis. 2011;17:1034–1043. [PMC free article] [PubMed] [Google Scholar]
  • 145.Ha Y, et al. Diabetes accelerates retinal ganglion cell dysfunction in mice lacking sigma receptor 1. Mol Vis. 2012;18:2860–2870. [PMC free article] [PubMed] [Google Scholar]
  • 146.Su TP, Hayashi T. Understanding the molecular mechanism of sigma-1 receptors: towards a hypothesis that sigma-1 receptors are intracellular amplifiers for signal transduction. Curr Med Chem. 2003;10(20):2073–2080. doi: 10.2174/0929867033456783. [DOI] [PubMed] [Google Scholar]
  • 147.Hayashi T, Maurice T, Su TP. Ca(2+) signaling via sigma(1)-receptors: novel regulatory mechanism affecting intracellular Ca(2+) concentration. J Pharmacol Exp Ther. 2000;293(3):788–798. [PubMed] [Google Scholar]
  • 148.Debonnel G, et al. Differential effects of sigma ligands on the N-methyl-D-aspartate response in the CA1 and CA3 regions of the dorsal hippocampus: effect of mossy fiber lesioning. Neuroscience. 1996;71(4):977–987. doi: 10.1016/0306-4522(96)80001-7. [DOI] [PubMed] [Google Scholar]
  • 149.Monnet FP, et al. N-methyl-D-aspartate-induced neuronal activation is selectively modulated by sigma receptors. Eur J Pharmacol. 1990;179(3):441–445. doi: 10.1016/0014-2999(90)90186-a. [DOI] [PubMed] [Google Scholar]
  • 150.Ishikawa M, 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(8):878–883. doi: 10.1016/j.biopsych.2007.04.001. [DOI] [PubMed] [Google Scholar]
  • 151.Ishikawa M, et al. High occupancy of σ1 receptors in the human brain after single oral administration of donepezil: a positron emission tomography study using [11C]SA4503. Int J Neuropsychopharmacol. 2009;12(8):1127–1131. doi: 10.1017/S1461145709990204. [DOI] [PubMed] [Google Scholar]
  • 152.Maurice T. Protection by sigma-1 receptor agonists is synergic with donepezil, but not with memantine, in a mouse model of amyloid-induced memory impairments. Behav Brain Res. 2016;296:270–278. doi: 10.1016/j.bbr.2015.09.020. [DOI] [PubMed] [Google Scholar]

RESOURCES