Next Article in Journal
Depolarization and Hyperexcitability of Cortical Motor Neurons after Spinal Cord Injury Associates with Reduced HCN Channel Activity
Next Article in Special Issue
Exercise-Induced Fibroblast Growth Factor-21: A Systematic Review and Meta-Analysis
Previous Article in Journal
Transcription Factor ZmNAC20 Improves Drought Resistance by Promoting Stomatal Closure and Activating Expression of Stress-Responsive Genes in Maize
Previous Article in Special Issue
Alzheimer’s Disease: An Updated Overview of Its Genetics
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Transient Receptor Potential (TRP) Channels in Pain, Neuropsychiatric Disorders, and Epilepsy

Department of Biomedical Sciences, School of Medicine, University of Missouri Kansas City, Kansas City, MO 64108, USA
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2023, 24(5), 4714; https://doi.org/10.3390/ijms24054714
Submission received: 16 January 2023 / Revised: 21 February 2023 / Accepted: 22 February 2023 / Published: 1 March 2023

Abstract

:
Pharmacomodulation of membrane channels is an essential topic in the study of physiological conditions and disease status. Transient receptor potential (TRP) channels are one such family of nonselective cation channels that have an important influence. In mammals, TRP channels consist of seven subfamilies with a total of twenty-eight members. Evidence shows that TRP channels mediate cation transduction in neuronal signaling, but the full implication and potential therapeutic applications of this are not entirely clear. In this review, we aim to highlight several TRP channels which have been shown to mediate pain sensation, neuropsychiatric disorders, and epilepsy. Recent findings suggest that TRPM (melastatin), TRPV (vanilloid), and TRPC (canonical) are of particular relevance to these phenomena. The research reviewed in this paper validates these TRP channels as potential targets of future clinical treatment and offers patients hope for more effective care.

1. Introduction

The last century of scientific development, especially in neurological sciences, has emphasized describing ion channels that play various roles in pathology. Accurate descriptions can open avenues for potential treatments, alongside enabling a deeper understanding of diseases themselves. Transient receptor potential (TRP) channels are one set of those discovered that have a wealth of potential. They are a collection of proteins that mainly function as nonselective cation channels [1]. First observed in a mutant strain of Drosophila, contemporary science has delineated 28 different TRP channels across multiple species [2,3]. Looking at genetic differences, the TRP superfamily can be separated into seven different subfamilies: TRPC (canonical); TRPV (vanilloid); TRPM (melastatin); TRPP (polycystin); TRPML (mucolipin); TRPA (ankyrin); and TRPN (NOMPC-like) [4]. Each TRP family channel consists of six transmembrane helical domains (TM1-TM6), N- and C-terminal regions in the cytosol, and a loop between the TM5 and TM6 domains, forming the important channel pore (Figure 1) [5]. Each TRP is being investigated as a possible treatment target for cerebrovascular disease, pain, psychological and neurological illnesses, diabetes, and even cancer [1,6,7,8,9,10]. In this paper, emphasis is placed on TRPM, TRPC, and TRPV, concerning their roles in pain, psychiatric diseases, and epilepsy (Table 1).
Melastatin homology regions (MHRs) distinguish TRPMs from the TRP superfamily, in addition to the fact that TRPMs have the most amino acids in their systolic domain. The TPRM family can be further subdivided into TRPM1-TRPM8, where sequence homology produces related pairs as follows: TRPM1 and TRPM; TRPM2 and TRPM8; TRPM4 and TRPM5; and the last pair, TRPM6 and TRPM7 [5]. All of the TRPMs permit different ions to pass through the channel pore with some sharing similar permeability and others varying. (For an in-depth look at ion permeability, see the cited review [11]). Each of the TRPMs has a conserved Ca2+-binding site, but recent data suggest that only TRPM2 and TRPM8 require it for gating [11,12,13]. TRPM4 and TRPM5 are uniquely the only TRPM channels impermeable to divalent ions such as Ca2+, even though these are Ca2+-activated channels [14]. Noted mutations in genes encoding TRPMs create channelopathies that influence cancer, neuropathic pain, inflammation, hypertension, diabetes, and hypomineralization [11].
TRPC channels were given their name—TRP canonical—due to the fact they most resembled Drosophila TRP channels that were found to be responsible for light sensing [15]. There are seven members of this subfamily, TRPC1-7, which can also be further subdivided into four groups based on their sequence homology: TRPC1, TRPC2, TRPC3/6/7, and TRPC 4/5 [16]. Other studies have noted that TRPC2 is found to be a pseudogene in humans and have thus organized the subfamilies into TRPC1/4/5 and TRPC3/6/7 [17]. Mammalian TRPC channels are activated downstream from receptors that signal using phospholipase C (PLC) but are expressed in numerous tissue types with various functions that differ from similar Drosophila proteins [18]. Due to the relationship between the Ca2+ filling store and PLC, these channels sometimes become activated in response to changing Ca2+ levels and are permeable to cations, namely Na+ and Ca2+ influx [18]. Research has shown that these channels play important roles in cardiovascular and renal health, as well as being potential targets for epilepsy treatment [19,20,21,22].
TRPV1 was the first mammalian TRP channel to have its structure delineated and to be cloned [23,24]. Of this subfamily, TRPV1-4 functions as the thermal sensing channels, which are temperature-activated, while TRPV5-6 functions are not activated by temperature and instead take the more conventional TRP route of being calcium-sensitive [25]. TRPV1, 2, and 4 have been shown to influence cerebral ischemic injury where these channels demonstrate a neuroprotective effect via Ca2+ influx and other signaling pathways, such as JNK and p38 MAPK [26]. Further research observed TRPV1 channel activation inhibiting neutrophil infiltration and reducing free radical generation in ischemia–reperfusion injury [27]. Together, these membrane subtypes pose significant opportunities for future research and biomedical innovation.
Table 1. TRP Channel subfamily is involved in pain, psychiatric, and epileptic disorders.
Table 1. TRP Channel subfamily is involved in pain, psychiatric, and epileptic disorders.
TRP Channel SubfamilyPain DisordersPsychiatric DisordersEpileptic Disorders
TRPM
  • TRPM2 KO reduces nocifensive response in inflammatory, neuropathic, mechanical, and thermal pain [28,29,30,31]
  • TRPM3 facilitates heat nociceptive sensation [26,32]
  • TRPM8 likely facilitates cold afferent sensation in context of analgesia [33,34]
  • TRPM2 KO precipitates bipolar behaviors [35,36]
  • TRPM3 implicated in post-partum mood disorders, such as depression and anxiety [37,38]
  • TRPM2 plays a protective role against JME-associated cell death in relation to EFHC1 [39]
  • TRPM3 overactivity associated with DEE [40,41]
  • TRPM7 inhibition reduces seizure-induced neuronal death [42,43]
TRPV
  • TRPV1 inhibition resulted in increased heat pain threshold and reduced osteoarthritic pain [44,45,46]
  • TRPV4 inhibition resulted in decreased neuropathic pain in chronic back pain models [47]
  • TRPV1 KO demonstrated less anxiety, freezing, and contextual fear behaviors [48,49,50,51]
  • Increased TRPV1 expression with methamphetamine use [52]
  • TRPV1 agonism with capsaicin demonstrated pro-convulsant activity [53]
  • TRPV1 antagonism with CPZ has protective role against neuronal apoptosis and epilepsy [54]
TRPC
  • TRPC3 likely signals both itch sensation and pain [55]
  • TRPC5 KO had increased inflammatory joint pain but decreased touch pain in sickle cell disease, migraine, chemotherapy-related pain, and surgical pain [56]
  • TRPC5 implicated in transferring conditioned fear responses to amygdala [57,58]
  • TRPC4 and 5 inhibition associated with antidepressant and anxiolytic behavioral phenotype [58,59]
  • Increased TRPC3 expression decreases threshold for epileptic activity [60]
  • Increased TRPC1 expression in FCD and seen in astrocyte modulation of epilepsy [61,62]
  • TRPC6 downregulated in chronic epileptic conditions [63]

2. TRPM

Table 1 has shown the TRPM in pain, psychiatric and epileptic disorders.

2.1. Pain Disorders

The basic mechanism of pain can be broken down into transduction, transmission, and modulation with the introduction of a noxious stimulus [64]. Furthermore, calcium ions play an important role in generating action potentials in order for nociceptive signals to be sent out through neurons [64]. Ion channels, such as voltage-gated calcium channels (VGCCs), have been well established as integral mediators of pain [65]. The role of calcium in nociception is highlighted in our review of TRP channels as several studies have portrayed this family of receptors as an important modulating target for analgesia in acute and chronic pain [66,67]. Although it has been speculated that TRPM ion channels have both pain-exacerbating and pain-alleviating mechanisms, the majority of studies using both agonism and antagonism of TRPM subtypes support their pro-inflammatory and pain-inducing functions [28,29].
TRPM2 channels are one such family of proteins that have interactions with reactive oxidative species (ROS) [30] and is involved in pathological pain [31]. The current understanding of the receptor is that it acts as a sensor for both ROS and adenosine diphosphate ribose (ADPR) in calcium gating [31]. Its expression in mice is mostly seen in primary afferent sensory neurons, including both A and C fiber neurons [31]. TRPM2 knock-out (KO) mice in a study conducted by Haraguchi et al. demonstrated reduced nocifensive response in inflammatory, neuropathic, mechanical, and thermal pain, without the loss of baseline sensitivity to mechanical and thermal sensation [32]. Primarily expressed in somatosensory dorsal root ganglion (DRG) neurons, thermogenic TRPM3 plays an integral role in heat nociception. TRPM3 knockout mice displayed a decreased response to noxious heat stimuli [26,68]. Blockage of the receptor by hesperetin, isosakuranetin, and primidone decreased intracellular transport of calcium ions, and similar to TRPM3 knockout mice, rodents treated with flavanones and primadone displayed decreased pain behaviors in response to noxious heat stimuli [33].
The development of the antagonists and modulators of TRPM channels is already underway, especially for TRPM8. TRPM8 was first characterized as a receptor for cold sensory afferents, which are activated by certain natural molecules known to produce this sensation, such as menthol, icilin, and eucalyptol [34]. The goal of cold sensation in the context of this receptor’s activation is the promotion of analgesia. However, TRPM8 is also believed to be involved in pain amplification due to cold hypersensitivity. While more research must be conducted to identify a definitive role in thermal-associated pain, knockout mice of this gene lacked cold hypersensitivity [69]. Other implications of pain for TRPM8 include migraines and bladder pain [70,71,72]. It is likely that TRPM8 is expressed for cold-afferent functions, and the modulation of the receptor shows potential for thermoreceptive pain analgesia.

2.2. Psychiatric Disorders

Unfortunately, most psychiatric illnesses are without complete descriptions of their etiology. One new theory contends that the oxidative state of cells and subsequent downstream calcium changes may have an influence [73]. While it is highly contested whether the antioxidant system is upregulated or downregulated in psychiatric disorders, many studies demonstrate changes in superoxide dismutases (SODs), catalase (CAT), and glutathione peroxidases (GPXs) that correlate with mental health issues [73,74]. A meta-analysis of double-blind, randomized, placebo-controlled trials evaluating the effects of N-acetyl cysteine (NAC) found that treatment significantly improved depressive symptoms in heavy smokers, those with trichotillomania, and those with bipolar or depressive disorders [73,75]. One double-blind, randomized, placebo-controlled, clinical trial of its efficacy as an adjunctive treatment for patients with schizophrenia on antipsychotic medications showed that patients taking NAC presented improvements on the Positive and Negative Syndrome Scale (PANSS) compared to the placebo group [76]. The same trial demonstrated significantly improved cognitive functions [76]. However, how exactly do TRPM channels relate to the psychiatric significance of redox reactions at the cellular level?
The TRPM2 channel is among a group of highly sensitive proteins that are influenced by redox status [73]. When they are activated, they induce Ca2+ entry and transfer redox activity into intracellular Ca2+ signaling [73]. Specific details have been more thoroughly reviewed elsewhere [77]. With only this information, one can see the theoretic bridge between redox reactions, intracellular calcium, and psychiatric illness that TRPM channels provide. Two studies have demonstrated the need for TRPM2 in order to have NMDAR-dependent long-term depression in the hippocampus [78,79]. Another investigation demonstrated the key role that TRPM2 channels play in the H2O2-dependent modulation of substantia nigra pars reticulata GABAergic neurons [35].
Fine mapping linkage analysis showed that TRPM2 on chromosome 21q is associated with bipolar disorder [36]. One of the identified single nucleotide polymorphisms (SNPs) leads to a deletion of TRPM2, and subsequent studies demonstrated that TRPM2 KO mice exhibit bipolar-disorder-related behaviors [37]. What is even more intriguing is that one of the essential targets of lithium is GSK-3, which experiences increased phosphorylation whenever TRPM2 is disrupted [37,79]. Other findings expand the likely effects of TRPM2 channels to include major depressive disorders, as well as other behavioral phenotypes that are related to oxidative stress [38]. Beyond TRPM2, recent research highlights TRPM3 as a potential player in the development of mood and anxiety disorders, with specific mention of post-partum mood disorders [80]. Similar efforts revealed that TRPM3 expression was changed in a mouse model of bipolar disorder, furthering the aforementioned evidence [81].

2.3. Epileptic Disorders

Epilepsy is a chronic neurological disorder caused by abnormal electrical excitability in the brain. Calcium ion (Ca2+) accumulation has been hypothesized to play a critical role in the etiology of this disease [82]. Though not fully understood, the activation of tissue transglutaminase (tTG) by calcium accumulation is thought to trigger glutamate-induced neurotoxicity that yields seizure activity [83]. This hypothesis is like that of neuronal injury in brain ischemia and trauma, where the amount of calcium influx correlates with the degree of damage [84]. Loss-of-function K+ channels and gain-of-function Na+ channels have also been identified as contributors to neuron excitability and causes of epilepsy. TRP channels are nonselective transmembrane cation proteins that allow cations such as Ca2+ to pass through their pores [39,85,86]. Because ionic disbalance plays a central role in the etiology of epilepsy, these nonselective TRP channels serve as areas of interest within the field.
Three TRPM channels that have shown direct effects in epilepsy include TRPM2, TRPM3, and TRPM7. TRPM2 is co-expressed and has direct physical interactions with the EF-hand motif-containing protein, EFHC1. Typically, EFHC1 enhances the susceptibility of TRPM2 to neuronal apoptosis through ROS and H2O2. In juvenile myoclonic epilepsy (JME), however, there is a mutation in EFHC1. Given the direct relationship between EFHC1 and TRPM2, it is suggested that TRPM2 plays a protective role against cell death that contributes to the phenotypic presentation of JME [40].
Developmental and epileptic encephalopathies (DEEs) are groups of conditions characterized by epilepsy with comorbid developmental delay. Gain-of-function mutations that cause overactivity in the TRPM3 channel are associated with DEEs [41]. Like most other TPR channels, TRPM3 is a nonselective cation channel that is permeable to calcium [42]. Primidone, a clinically approved antiepileptic, is a TRPM3 antagonist whose antiepileptic effects could be in part due to its inhibition of TRPM3 [41]. Therefore, the development of TRPM3 inhibitors serves as a focus for future epileptic therapeutics.
Within the brain, TRPM7 is activated and expressed during epilepsy, perpetuating a positive feedback loop and contributing to the production of ROS. This is confirmed by the fact that TRPM7 ablation prevents a cation current under circumstances of oxygen–glucose deprivation, and therefore prevents ROS-mediated cell death [43]. A recent study published in the International Journal of Molecular Sciences demonstrated that TRPM7 inhibition reduced the seizure-induced expression of TRPM7 channels [87]. This resulted in decreased intracellular zinc accumulation, ROS production, and postictal apoptotic neuronal death [87].
Other TRPM channels have demonstrated regulatory roles that could translate to the pathogenesis of epilepsy if found within epileptic regions of the brain. GTL-2 is a TRPM that influences local ion composition in non-neuronal tissues. Its function in the epidermis has been compared to the modulatory role of glial cells, as GTL-2 channels contribute to ionic buffering and glutamate uptake that regulate neurotransmitter release [88]. If found in neuronal tissues, GTL-2 could play a role in the ionic modulation of Ca2+ that yields glutamate release during seizures. In the mouse retina model, TRPM1 localized to the dendrites of ON-bipolar cells and modulated depolarization in response to light-induced glutamate release [89]. Similar to GTL-2, if TRPM1 contributes to depolarization and glutamate release in brain regions impacted during epileptic episodes, its inhibition could hold a significant role in regulating seizures.

3. TRPV

Table 1 also has shown the TRPV in pain, psychiatric and epileptic disorders.

3.1. Pain Disorders

Among the TRP channels, ones belonging to the TRPV family appear to have the strongest backing from the literature, supporting their roles in modulating pain sensation. TRPV1 is expressed in a variety of tissues, including sensory ganglia and small sensory neural fibers [90]. More than a decade ago, early rodent models using TRPV1 antagonism highlighted the receptor’s role in thermosensation and pain perception, similar to that of TRPM [44]. This idea was further tested in human trials, where TRPV1 antagonist SB-705498 produced significant results in increasing the heat pain threshold in treated skin [45].
More recently, one study reported inhibition of pain behavior in osteoarthritis (OA) mice after intra-articular administration of JNJ-17203212, a TRPV1 antagonist [46]. This was measured by evaluating the attenuation of weight-bearing asymmetry after antagonism of the receptor [46]. The same study also found that there was an increased expression of this receptor in human OA, rheumatoid arthritis (RA), and postmortem (PM) synovium via immunohistochemistry [46]. Based on this evidence, modulation of TRPV1 may prove to be beneficial in attenuating inflammatory joint pain, especially because TRPV1 has been seen to respond to oral paracetamol antinociception in human patients [91]. In addition to joint pain, TRPV1 expression in microglia in anterior cingulate cortices, GABAergic spinal interneurons, as well as other sensory ganglia, plays roles in numerous pain mechanisms of neuroinflammation, allodynia, and mechanical hyperalgesia [92,93,94]. Interestingly, cannabidiol (CBD) is another molecule that has been tested to modulate TRPV1 to observe its role in analgesic pathways. Cannabis-based products have become easily accessible and clinical trials in humans have supported their usage in self-medicated pain relief [29,95]. In a Parkinson’s disease mouse model, CBD increased anandamide binding to TRPV1 receptors and produced analgesic effects [96]. CBD’s exertion of pain alleviation may be dose-dependent as one group found that concentrations of 10 and 50 μmol/L CBD were required for decreased calcium shuttling in TRPV1 channels [47].
The TRPV4 receptor is expressed in a variety of tissues, including immune cells, sensory neurons, glial cells, the spinal cord, cortical pyramidal neurons, the thalamus, and cerebellum basal nuclei [97]. Similar to TRPV1, TRPV4 is thermosensitive and can be activated by heat, as well as by mechanical forces [97]. One of the most important roles of TRPV4 found in studies is the alleviation of neuropathic pain via its inhibition. A major source of chronic pain commonly seen in patients is chronic back pain. In several studies, TRPV4 antagonism was seen to reduce cartilage degradation after mechanical compression, reduce IL-1β-mediated NO release, and decrease neuropathic pain in chronic compression of DRG pain models in rats [98,99,100]. One pre-clinical candidate thought to inhibit TRPV4, named GSK3527497, was described to be a potential TRPV4 antagonist that can be used for therapeutic pain management [101]. More recently, GSK3527497 generated favorable results and was found to be well tolerated in Phase I clinical trials in both healthy volunteers and heart failure patients after 14 days of dosing [57]. While more studies utilizing human subjects are needed, these findings suggest antagonists of TRPV4 could be a new type of pharmacotherapy in the future.

3.2. Psychiatric Disorders

As mentioned in the section discussing TRPMs, TRP channels are suggested to play a role in the etiology of psychiatric illnesses. In the TRPV family, TRPV1 has amassed the most evidence, indicating its relevance in the realm of mental health [48]. For example, TRPV1 KO mice have shown less anxiety-related behavior, freezing, and contextual fear in different methodological designs [49]. These findings were paralleled by an impairment in hippocampus-dependent learning, consisting of a deficit in long-term potentiation, not unlike the TRPM2 findings [49]. Further investigations have examined the role of these physiological manipulations in the etiologies of mental disorders.
One study looked at the antidepressant effects of TRPV1 agonists and found that the administration of capsaicin and olvanil provided significant protective effects against induced depression in mice [50]. Another experiment focusing on the ventral hippocampus in rats found that TRPV1 channels have an important role in regulating anxiety [51]. This was shown via the antagonism of the TRPV1 channels via capsazepine and was redemonstrated by Kasckow et al. [50]. Outside of the hippocampus, similar findings have been observed in the periaqueductal grey where, again, antagonism of TRPV1 led to anxiolytic-like effects [52]. Another set of mental health issues, namely substance use disorders, are also found to correlate with TRPV1 activity. Multiple studies demonstrate that continued methamphetamine exposure leads to increases in TRPV1 expression, specifically in the frontal cortex [102]. Another investigation found that the deletion of TRPV1 channels led to an altered behavioral response to ethanol administration in mice [103]. Together, these findings demonstrate that TRPV1 is a significant target for future research regarding the treatment of psychiatric diseases.

3.3. Epileptic Disorders

TRPV1 is one of the most well-characterized and documented transient-receptor potential-vanilloid channels in the TRPV family. Like other TRP channels, it yields nonselective Ca2+ permeability. TRPV1 was initially found to be highly expressed in the DRG, the sensory neurons. This paved the way for the discovery of the significant role of TRPV1 in inflammatory hyperalgesia [24]. Subsequent research demonstrated that TRPV1 is also located in various nuclei locations within the human brain, including the cortex, hypothalamus, and hippocampus [53]. Glutamate excitation through Ca2+ activation, especially in the calcium-sensitive hippocampus, has long been an important etiologic consideration for epilepsy [39]. The function of TRPV1 in hippocampal excitation thereby highlights a new focus in the field.
Capsaicin is a TRPV1 agonist that can be used to elicit the effects of the channel. In the mouse model, capsaicin demonstrated pro-convulsant activity that was only blocked by pretreatment with capsazepine (CPZ), an antagonist of TRPV1 [104]. Further, systemic or hippocampal administration of a TRPV1 antagonist reduced the susceptibility of mice to pentylenetetrazol (PTZ)-induced seizures [54]. The activation of TRPV1 by capsaicin caused apoptosis in the hippocampi and DRG of rats, indicating that this channel is important for regulating epileptic episodes. Conversely, TRPV1 blockers such as CPZ demonstrated a protective role against neuronal apoptosis and epilepsy in the same rat model [105]. Cannabidiol (CBD) is argued to be an antagonist and agonist of TRPV1 and TRPV2, respectively [47,106]. In either case, CBD administration decreased in vitro epileptiform activity and in vivo seizure activity in rats [107]. In a descriptive study of humans, increased expression of TPRV1 mRNA/proteins was found in patients with mesial temporal lobe epilepsy compared to the control [55]. These findings provide evidence of the role of TRPV1 functions in perpetuating seizures among animals and likely humans. Future work concentrating on TRPV1 inhibitors as antiepileptic therapeutics remains a promising area to be researched.

4. TRPC

Table 1 has shown the TRPC in pain, psychiatric and epileptic disorders.

4.1. Pain Disorders

All TRPC channels reveal calcium permeability, which is an important part of pain transduction and sensitization of nociceptors as previously mentioned. Based on this cellular mechanism, it is worthwhile to continue the investigation of the TRPC channels’ role. While new research surrounding TRPC3 leans more toward its role in non-histaminergic itch, it is likely that TRPC3 signals both itch and pain, given that it is expressed heterogeneously in both nociceptors and pruriceptors [108]. In situ hybridization in one study found that TRPC3 was exclusively expressed in small-to-medium-diameter sensory neurons in rat DRG [109,110]. Disruption of endogenous TRPC3 in these ganglia resulted in decreased expression of store-operated calcium entry (SOCE), UTP, or protease-activated receptor 2 (PAR2) agonist-evoked calcium transduction [109]. TRPC3′s contribution to calcium-induced nociceptor sensitization and its relationship with co-expression with other proinflammatory receptors support its importance in future pain therapies. The strongest support for the future of TRPC-related non-opioid analgesia comes from research surrounding TRPC5. Recently, the function of this receptor in pain perception was observed in murine models of osteoarthritis where TRPC5 KO mice had significantly exacerbated pain-like behaviors compared to that of the wild type (WT) [111]. Along with behavioral results, findings were also found at the microscopic level where TRPC5 KO mice had increased mast cell markers and extracellular matrix remodeling, synonymous with synovial inflammation [111]. One group shared similar findings with TRPC5 gene-deleted mice and saw increased weight-bearing asymmetry, inflammatory cytokines, and secondary hyperalgesia in human inflammatory arthritis synovia after chronic treatment of a TRPC5 antagonist (ML204) [56]. While some osteoarthritis models suggest TRPC5 inhibition leads to more inflammation and pain, other researchers saw reversed touch pain in mice models of sickle cell disease, migraine, chemotherapy-related pain, and surgical pain after TRPC5 inhibition [58]. TRPC5 appears to have great translation potential as well, due to its high expression in donor human DRG tissues [58].

4.2. Psychiatric Disorders

Of all the TRP channels, TRPC channels have become the biggest target for potential treatments [48]. For one, TRPC5 is thought to be responsible for transferring conditioned fear responses to the amygdala, as they are expressed in the thalamus, amygdala, and cortical areas related to fear responses [48,112]. TRPC4 was identified by the same group of researchers to be essential for behavioral responses to anxiety-inducing stimuli, seeing as KO mice demonstrated an anxiolytic behavioral phenotype [112]. Because TRPCs are expressed in B lymphocytes, one study found that expression on those cells was changed in those with bipolar disorder, suggesting a functional change in the channel in the disease state [113]. Interestingly, treatment for 7 days with lithium reduced the TRPC3 protein in those same cells in bipolar patients [114].
Looking more at the impact on neuronal populations implicated in the reward system and other goal-oriented behavior, one study found that TRPC1 deletion led to a loss in striatal cells, in addition to proteomic alterations [115]. A further physiological mystery of the TRPC channels in the brain was found in an investigation of IL-10 KO mice. These mice had reduced TRPC5 expression in the medial prefrontal cortex and amygdala [59]. These mice then demonstrated enhanced depressive and anxiety-like behavior, potentially relating the immune system to the TRPC channels already implicated in depression and anxiety [59]. Some researchers have taken these TRPC findings far enough to develop treatments with novel agents.
In fact, acute treatment with a novel TRPC4/5 inhibitor was found to produce antidepressant and anxiolytic effects in mice [116]. Furthermore, they posit that these effects are mediated through downstream signaling involving BDNF [116]. Another study with a different but equally novel TRPC4/5 inhibitor was able to generate similar findings [60]. The compound reduces CCK-4-invoked neuronal activity in amygdala slices, the brain region known to produce fear responses [60]. Together, these findings suggest once more that TRP channels are implicated in psychiatric illness and physiology, specifically the TRPC channels mentioned above.

4.3. Epileptic Disorders

Among the TRPC families, TRPC3 is the most reviewed within the setting of epilepsy. While all TRP channels regulate cations in some regard, TRPC3 uniquely mediates low Mg2+ and Ca2+ depolarization contributing to epilepsy. In addition, higher expression of TRPC3 makes the immature cortex more excitable and thus decreases the threshold for epileptic susceptibility [117]. In animal models of pilocarpine-induced status epilepticus, TRPC3 has been studied using different pharmacologic agents. The main outcomes from targeting this channel with Pyr3, a TRPC3 inhibitor, included decreases in TRPC3 channel expression and reductions in the root mean square of power/theta activity of seizures [118]. In hyperthermia-induced febrile seizure rats, elevated mRNA and protein levels were detected in neuronal cells within the hippocampus. Accordingly, when the Pyr3 inhibitor was administered in this model, seizure severity, neuroinflammation, and neuronal cell death decreased [119]. While Pyr3 is a selective and potent TRPC3 inhibitor, its metabolic instability and toxicity limit its full benefit. A modified pyrazole compound, JW-65, was recently studied for its effects on TRPC3. Systemic administration of JW-65 prior to pilocarpine induction showcased a delay in seizure initiation. JW-65 administration after pilocarpine-induced seizures initiated anti-seizure activity, confirmed by electroencephalographic monitoring with video. Seizure susceptibility decreased in a dose-dependent manner in accordance with JW-65 administration [61]. These findings suggest that JW-65 is a novel therapeutic of interest in seizure attenuation. In conjunction, the results of these TRPC3 studies demonstrate that TRPC3 is important in promoting electrical excitability within various seizure models. Inhibition of this channel proves significant to reducing seizure activity. Thus, TRPC3 channel inhibitors serve as an area of interest in the field of anticonvulsant therapeutics.
TRPC1 and TRPC6 are two other TRPC channels of epileptogenic interest [62]. Among humans with focal cortical dysplasia (FCD), increased mRNA and protein expressions of TRPC1 were found in cortical tissue. Such findings of overexpression in cell-specific patterns suggest that TRPC1 may play a role in creating conditions permissive for FCD. Notably, TRPC1 also showed colocalization with GFAP in reactive astrocytes, indicating that the channel may aid in astrocyte modulation of epilepsy [63].
TRPC6 holds unique consideration in epilepsy, as it was shown to be downregulated under chronic epileptic conditions in the rat model. Furthermore, ablation of TRPC6 by protein-specific silencing RNA causes increased susceptibility to pilocarpine-induced seizures. This may be because TRPC6 is involved with signaling the expression of proteins important to mitochondrial function [120]. A deeper understanding of TRPC6 in neuroprotective signaling pathways will be beneficial for understanding the impact of this channel, especially in the setting of epilepsy.
TRP channel subfamily involved in pain, psychiatric, and epileptic disorders have been shown in Table 1.

5. Conclusions

TRP channels are a group of multifunctional membrane proteins that serve important functions in neurological physiology and pathology. They significantly influence oxidative species sensing, afferent nociceptive signaling, bipolar disease, depression, anxiety, and seizure disorder [35,39,66,67,78,121,122]. In addition, TRP channels play important roles in cardiovascular and renal health [19,20]. Most TRPs highlighted in this review are subjects of pharmacological development and intense study, especially to mitigate the progression of these pathological conditions. The aim of this review was to examine TRPM, TRPV, and TRPC channels and gain perspective and understanding of their growing impact on pain sensation, psychiatric diseases, and epilepsy.
Regarding pain, TRP channels represent an area of focus that is of high interest to researchers and pharmaceutical companies to develop non-opioid interventions for chronic pain. Currently, medical management of chronic pain remains dominated mainly by opioid analgesics, but the downsides of addiction and tolerance make it an unsatisfactory long-term treatment option for pain [123,124]. Thus, advancements in modulating TRP channels and nociceptive signaling from the peripheral nervous system are a more favorable route. TRPM8 is implicated in cold hypersensitivity and is theorized to alleviate inflammatory or neuropathic pain, but there is still controversy regarding the effects of TRPM8 agonism and antagonism on pain perception [34,69,125]. Inhibition of TRPV1 and TRPV4 is related to several pain-alleviating findings, especially neuropathic pain and mechanical hyperalgesia seen in inflammatory joint disease [46,91,92,93,94]. TRPV1 is also believed to serve a first-line defense role in noxious heat stimuli via capsaicin [126]. A growing field of pharmacological therapeutics that target TRPV1 includes CBD, which decreases Ca2+ transduction in DRG after stimulation with capsaicin [29,47,95]. In the setting of chronic neuropathic pain, non-opioid antagonism of TRPV4 shows promising data to be a major analgesic target [57]. While TRPC3 shares similar characteristics to the previously mentioned properties of TRPs, TRPC5′s physiological role likely leans more toward reducing inflammation and pain [56,108,111].
Although several psychiatric diseases do not have completely clear mechanisms of pathology, a wealth of research support TRP channel expression and function in the setting of neuropsychiatric diseases. In fine-mapping linkage of bipolar subjects, TRPM on chromosome 21q is seen to be associated with bipolar disease behaviors [36,37]. While lithium is traditionally used for bipolar symptoms, TRPM2 is a candidate for several new compounds developed within the last 5 years. Namely, ADPR analog TRPM2 antagonists may prove beneficial in the future for a spectrum of neuropsychiatric diseases, including bipolar disease [127]. There is literature supporting TRPV1′s antagonism benefiting anxiety, and its agonism promoting protection against depression, anxiety, and substance use disorder [49,51,102,103,128,129]. Antagonism of TRPV1 is a possible route for the reduction in anxiety and fear responses with future translational research [49,129]. Continuing the mechanisms of anxiety, TRPC4 and TRPC5 are heavily implicated in the transmission of conditioned fear responses to the amygdala, and gene disruption in mice demonstrated anxiolytic effects [48,112]. Two antagonists of TRPC4/5, M084 and HC-070, have already been developed, and researchers found that the administration of these two agents induced anxiolytic and anti-depressant effects in mice [60,116]. Together, these findings highlight TRPC’s important role in the physiology of anxiety and depression.
Cation transduction, including that of calcium ions, is believed to play a massive role in the pathology of epilepsy and seizures, possibly due to the glutamate-induced neurotoxicity from tissue transglutaminase activation [82,83,130]. Because ion imbalances serve a central role in the etiology of epilepsy, TRP channels are a major area of interest in understanding and treating epilepsy [82,85,86]. TRPM2 may be a channel protein of interest in juvenile myoclonic epilepsy as EFHC1, the candidate gene for this disease, potentiates TRPM2 in ROS-mediated neuronal death [40]. Primidone, along with diclofenac and maprotiline, have all been demonstrated to be efficient blockers of TRPM3 with IC50 at 0.6–6 μM [131]. Other TRPM channels that have the potential to be therapeutic targets in epilepsy include TRPM7, where inhibition reduced the seizure-induced expression of TRPM7 channels and reduced ROS-related neuronal death [43]. Agonism of TRPV1, which is expressed in the brain, hypothalamus, and hippocampus, with capsaicin and pentylenetetrazol (PTZ) demonstrated pro-convulsant activity that was only blocked with TRPV1 antagonism [54,104,126,132]. Microscopically, researchers saw apoptotic mechanisms in hippocampal and DRG cells of rats after capsaicin TRPV1 agonism, which strengthens its relationship with seizure etiology [105]. The administration of CBD also decreased in vitro and in vivo seizure activity in rat [107]. Its direct effects on TRPV1 and TRPV2 is not certain as it more strongly modulates cannabinoid receptors. Expression of several TRPC channels is consistently seen to have an impact on chronic epileptic conditions, and TRPC3 uniquely mediates low Mg2+ and Ca2+ depolarization, contributing to epilepsy [61,62,63,117,118,120].
There is abundant information and research highlighting TRP channels and their roles in pathological processes involving cation transduction through cellular membranes. Clinical trials regarding TRP modulation are already undergoing and have generated promising results. Future investigations into pharmacological interventions of TRP channels will propagate further developments in patient treatments, especially in the setting of pain, neuropsychiatric diseases, and epilepsy.

Author Contributions

Conceptualization, supervision, methodology, project administration, formal analysis, investigation, funding acquisition, validation, writing—review and editing, X.-P.C.; Methodology, formal analysis, investigation, writing—original draft preparation, formal analysis, F.Y., A.S., V.C. and S.S. All authors have read and agreed to the published version of the manuscript.

Funding

This work was partly supported by the American Heart Association (19AIREA34470007), USA (X.P.C.).

Institutional Review Board Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We would like to thank the University of Missouri-Kansas City School of Medicine student research program for their support of S.S., A.S., V.C. and F.Y.’s professional studies. F.Y., A.S., S.S. and V.C. are recipients of the Sarah Morrison Student Research Award from the University of Missouri-Kansas City School of Medicine. X.-P.C. acknowledges the support from the American Heart Association (grant number: 19AIREA34470007). We would also like to thank BioRender for its application in developing the figures and table.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Fallah, H.P.; Ahuja, E.; Lin, H.; Qi, J.; He, Q.; Gao, S.; An, H.; Zhang, J.; Xie, Y.; Liang, D. A Review on the Role of TRP Channels and Their Potential as Drug Targets_An Insight into the TRP Channel Drug Discovery Methodologies. Front. Pharmacol. 2022, 13, 914499. [Google Scholar] [CrossRef] [PubMed]
  2. Cosens, D.J.; Manning, A. Abnormal Electroretinogram from a Drosophila Mutant. Nature 1969, 224, 285–287. [Google Scholar] [CrossRef] [PubMed]
  3. Samanta, A.; Hughes, T.E.T.; Moiseenkova-Bell, V.Y. Transient Receptor Potential (TRP) Channels. Subcell. Biochem. 2018, 87, 141–165. [Google Scholar] [CrossRef] [PubMed]
  4. Wang, R.; Tu, S.; Zhang, J.; Shao, A. Roles of TRP Channels in Neurological Diseases. Oxidative Med. Cell. Longev. 2020, 2020, 7289194. [Google Scholar] [CrossRef] [PubMed]
  5. Huang, Y.; Fliegert, R.; Guse, A.H.; Lü, W.; Du, J. A Structural Overview of the Ion Channels of the TRPM Family. Cell Calcium 2020, 85, 102111. [Google Scholar] [CrossRef] [PubMed]
  6. Zhang, L.; Papadopoulos, P.; Hamel, E. Endothelial TRPV4 Channels Mediate Dilation of Cerebral Arteries: Impairment and Recovery in Cerebrovascular Pathologies Related to Alzheimer’s Disease. Br. J. Pharmacol. 2013, 170, 661–670. [Google Scholar] [CrossRef] [Green Version]
  7. Takayama, Y.; Derouiche, S.; Maruyama, K.; Tominaga, M. Emerging Perspectives on Pain Management by Modulation of TRP Channels and ANO1. Int. J. Mol. Sci. 2019, 20, 3411. [Google Scholar] [CrossRef] [Green Version]
  8. García-Rodríguez, C.; Bravo-Tobar, I.D.; Duarte, Y.; Barrio, L.C.; Sáez, J.C. Contribution of Non-Selective Membrane Channels and Receptors in Epilepsy. Pharmacol. Ther. 2022, 231, 107980. [Google Scholar] [CrossRef]
  9. Derbenev, A.V.; Zsombok, A. Potential Therapeutic Value of TRPV1 and TRPA1 in Diabetes Mellitus and Obesity. Semin. Immunopathol. 2016, 38, 397–406. [Google Scholar] [CrossRef]
  10. Morelli, M.B.; Amantini, C. Transient Receptor Potential (TRP) Channels: Markers and Therapeutic Targets for Cancer? Biomolecules 2022, 12, 547. [Google Scholar] [CrossRef]
  11. Souza Bomfim, G.H.; Niemeyer, B.A.; Lacruz, R.S.; Lis, A. On the Connections between TRPM Channels and SOCE. Cells 2022, 11, 1190. [Google Scholar] [CrossRef] [PubMed]
  12. Zhang, Z.; Tóth, B.; Szollosi, A.; Chen, J.; Csanády, L. Structure of a TRPM2 Channel in Complex with Ca2+ Explains Unique Gating Regulation. eLife 2018, 7, e36409. [Google Scholar] [CrossRef]
  13. Yin, Y.; Le, S.C.; Hsu, A.L.; Borgnia, M.J.; Yang, H.; Lee, S.-Y. Structural Basis of Cooling Agent and Lipid Sensing by the Cold-Activated TRPM8 Channel. Science 2019, 363, eaav9334. [Google Scholar] [CrossRef] [PubMed]
  14. Winkler, P.A.; Huang, Y.; Sun, W.; Du, J.; Lü, W. Electron Cryo-Microscopy Structure of a Human TRPM4 Channel. Nature 2017, 552, 200–204. [Google Scholar] [CrossRef]
  15. Montell, C.; Birnbaumer, L.; Flockerzi, V.; Bindels, R.J.; Bruford, E.A.; Caterina, M.J.; Clapham, D.E.; Harteneck, C.; Heller, S.; Julius, D.; et al. A unifed nomenclature for the superfamily of TRP cation channels. Mol. Cell 2002, 9, 229–231. [Google Scholar] [CrossRef] [PubMed]
  16. Huang, J.; Du, W.; Yao, H.; Wang, Y. TRPC Channels in Neuronal Survival. In TRP Channels; Zhu, M.X., Ed.; CRC Press/Taylor & Francis: Boca Raton, FL, USA, 2011; ISBN 978-1-4398-1860-2. [Google Scholar]
  17. Liu, X.; Yao, X.; Tsang, S.Y. Post-Translational Modification and Natural Mutation of TRPC Channels. Cells 2020, 9, 135. [Google Scholar] [CrossRef] [Green Version]
  18. Wang, H.; Cheng, X.; Tian, J.; Xiao, Y.; Tian, T.; Xu, F.; Hong, X.; Zhu, M.X. TRPC Channels: Structure, Function, Regulation and Recent Advances in Small Molecular Probes. Pharmacol. Ther. 2020, 209, 107497. [Google Scholar] [CrossRef]
  19. Martín-Bórnez, M.; Galeano-Otero, I.; Del Toro, R.; Smani, T. TRPC and TRPV Channels’ Role in Vascular Remodeling and Disease. Int. J. Mol. Sci. 2020, 21, 6125. [Google Scholar] [CrossRef]
  20. Hall, G.; Wang, L.; Spurney, R.F. TRPC Channels in Proteinuric Kidney Diseases. Cells 2019, 9, 44. [Google Scholar] [CrossRef] [Green Version]
  21. Wen, H.; Gwathmey, J.K.; Xie, L.-H. Role of Transient Receptor Potential Canonical Channels in Heart Physiology and Pathophysiology. Front. Cardiovasc. Med. 2020, 7, 24. [Google Scholar] [CrossRef]
  22. TRPC Channels as Emerging Targets for Seizure Disorders: Trends in Pharmacological Sciences. Available online: https://www.cell.com/trends/pharmacological-sciences/fulltext/S0165-6147(22)00134-1#t0010 (accessed on 10 November 2022).
  23. Cao, E.; Liao, M.; Cheng, Y.; Julius, D. TRPV1 Structures in Distinct Conformations Reveal Activation Mechanisms. Nature 2013, 504, 113–118. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Caterina, M.J.; Schumacher, M.A.; Tominaga, M.; Rosen, T.A.; Levine, J.D.; Julius, D. The Capsaicin Receptor: A Heat-Activated Ion Channel in the Pain Pathway. Nature 1997, 389, 816–824. [Google Scholar] [CrossRef] [PubMed]
  25. Yelshanskaya, M.V.; Sobolevsky, A.I. Ligand-Binding Sites in Vanilloid-Subtype TRP Channels. Front. Pharmacol. 2022, 13, 900623. [Google Scholar] [CrossRef]
  26. Xie, Q.; Ma, R.; Li, H.; Wang, J.; Guo, X.; Chen, H. Advancement in Research on the Role of the Transient Receptor Potential Vanilloid Channel in Cerebral Ischemic Injury (Review). Exp. Ther. Med. 2021, 22, 881. [Google Scholar] [CrossRef] [PubMed]
  27. Randhawa, P.K.; Jaggi, A.S. A Review on Potential Involvement of TRPV1 Channels in Ischemia-Reperfusion Injury. J. Cardiovasc. Pharmacol. Ther. 2018, 23, 38–45. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. De Caro, C.; Cristiano, C.; Avagliano, C.; Bertamino, A.; Ostacolo, C.; Campiglia, P.; Gomez-Monterrey, I.; La Rana, G.; Gualillo, O.; Calignano, A.; et al. Characterization of New TRPM8 Modulators in Pain Perception. Int. J. Mol. Sci. 2019, 20, 5544. [Google Scholar] [CrossRef] [Green Version]
  29. Xu, D.H.; Cullen, B.D.; Tang, M.; Fang, Y. The Effectiveness of Topical Cannabidiol Oil in Symptomatic Relief of Peripheral Neuropathy of the Lower Extremities. Curr. Pharm. Biotechnol. 2020, 21, 390–402. [Google Scholar] [CrossRef]
  30. Kashio, M.; Sokabe, T.; Shintaku, K.; Uematsu, T.; Fukuta, N.; Kobayashi, N.; Mori, Y.; Tominaga, M. Redox signal-mediated sensitization of transient receptor potential melastatin 2 (TRPM2) to temperature affects macrophage functions. Proc. Natl. Acad. Sci. USA 2012, 109, 6745–6750. [Google Scholar] [CrossRef] [Green Version]
  31. Jang, Y.; Cho, P.S.; Yang, Y.D.; Hwang, S.W. Nociceptive Roles of TRPM2 Ion Channel in Pathologic Pain. Mol. Neurobiol. 2018, 55, 6589–6600. [Google Scholar] [CrossRef]
  32. Haraguchi, K.; Kawamoto, A.; Isami, K.; Maeda, S.; Kusano, A.; Asakura, K.; Shirakawa, H.; Mori, Y.; Nakagawa, T.; Kaneko, S. TRPM2 Contributes to Inflammatory and Neuropathic Pain through the Aggravation of Pronociceptive Inflammatory Responses in Mice. J. Neurosci. 2012, 32, 3931–3941. [Google Scholar] [CrossRef] [Green Version]
  33. Straub, I.; Krügel, U.; Mohr, F.; Teichert, J.; Rizun, O.; Konrad, M.; Oberwinkler, J.; Schaefer, M. Flavanones That Selectively Inhibit TRPM3 Attenuate Thermal Nociception in Vivo. Mol. Pharmacol. 2013, 84, 736–750. [Google Scholar] [CrossRef] [Green Version]
  34. McKemy, D.D.; Neuhausser, W.M.; Julius, D. Identification of a cold receptor reveals a general role for TRP channels in thermosensation. Nature 2002, 416, 52–58. [Google Scholar] [CrossRef]
  35. Lee, C.R.; Machold, R.P.; Witkovsky, P.; Rice, M.E. TRPM2 Channels Are Required for NMDA-Induced Burst Firing and Contribute to H(2)O(2)-Dependent Modulation in Substantia Nigra Pars Reticulata GABAergic Neurons. J. Neurosci. 2013, 33, 1157–1168. [Google Scholar] [CrossRef] [Green Version]
  36. McQuillin, A.; Bass, N.J.; Kalsi, G.; Lawrence, J.; Puri, V.; Choudhury, K.; Detera-Wadleigh, S.D.; Curtis, D.; Gurling, H.M.D. Fine Mapping of a Susceptibility Locus for Bipolar and Genetically Related Unipolar Affective Disorders, to a Region Containing the C21ORF29 and TRPM2 Genes on Chromosome 21q22.3. Mol. Psychiatry 2006, 11, 134–142. [Google Scholar] [CrossRef] [Green Version]
  37. Jang, Y.; Lee, S.H.; Lee, B.; Jung, S.; Khalid, A.; Uchida, K.; Tominaga, M.; Jeon, D.; Oh, U. TRPM2, a Susceptibility Gene for Bipolar Disorder, Regulates Glycogen Synthase Kinase-3 Activity in the Brain. J. Neurosci. 2015, 35, 11811–11823. [Google Scholar] [CrossRef]
  38. Andoh, C.; Nishitani, N.; Hashimoto, E.; Nagai, Y.; Takao, K.; Miyakawa, T.; Nakagawa, T.; Mori, Y.; Nagayasu, K.; Shirakawa, H.; et al. TRPM2 Confers Susceptibility to Social Stress but Is Essential for Behavioral Flexibility. Brain Res. 2019, 1704, 68–77. [Google Scholar] [CrossRef] [Green Version]
  39. Nazıroğlu, M. TRPV1 Channel: A Potential Drug Target for Treating Epilepsy. Curr. Neuropharmacol. 2015, 13, 239–247. [Google Scholar] [CrossRef] [Green Version]
  40. Katano, M.; Numata, T.; Aguan, K.; Hara, Y.; Kiyonaka, S.; Yamamoto, S.; Miki, T.; Sawamura, S.; Suzuki, T.; Yamakawa, K.; et al. The Juvenile Myoclonic Epilepsy-Related Protein EFHC1 Interacts with the Redox-Sensitive TRPM2 Channel Linked to Cell Death. Cell Calcium 2012, 51, 179–185. [Google Scholar] [CrossRef] [Green Version]
  41. Zhao, S.; Rohacs, T. The Newest TRP Channelopathy: Gain of Function TRPM3 Mutations Cause Epilepsy and Intellectual Disability. Channels 2021, 15, 386–397. [Google Scholar] [CrossRef]
  42. Grimm, C.; Kraft, R.; Sauerbruch, S.; Schultz, G.; Harteneck, C. Molecular and Functional Characterization of the Melastatin-Related Cation Channel TRPM3. J. Biol. Chem. 2003, 278, 21493–21501. [Google Scholar] [CrossRef] [Green Version]
  43. Aarts, M.M.; Tymianski, M. TRPMs and Neuronal Cell Death. Pflüg. Arch. 2005, 451, 243–249. [Google Scholar] [CrossRef]
  44. McGaraughty, S.; Chu, K.L.; Faltynek, C.R.; Jarvis, M.F. Systemic and Site-Specific Effects of A-425619, a Selective TRPV1 Receptor Antagonist, on Wide Dynamic Range Neurons in CFA-Treated and Uninjured Rats. J. Neurophysiol. 2006, 95, 18–25. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Chizh, B.A.; O’Donnell, M.B.; Napolitano, A.; Wang, J.; Brooke, A.C.; Aylott, M.C.; Bullman, J.N.; Gray, E.J.; Lai, R.Y.; Williams, P.M.; et al. The Effects of the TRPV1 Antagonist SB-705498 on TRPV1 Receptor-Mediated Activity and Inflammatory Hyperalgesia in Humans. Pain 2007, 132, 132–141. [Google Scholar] [CrossRef] [PubMed]
  46. Kelly, S.; Chapman, R.J.; Woodhams, S.; Sagar, D.R.; Turner, J.; Burston, J.J.; Bullock, C.; Paton, K.; Huang, J.; Wong, A.; et al. Increased Function of Pronociceptive TRPV1 at the Level of the Joint in a Rat Model of Osteoarthritis Pain. Ann. Rheum. Dis. 2015, 74, 252–259. [Google Scholar] [CrossRef] [PubMed]
  47. Anand, U.; Jones, B.; Korchev, Y.; Bloom, S.R.; Pacchetti, B.; Anand, P.; Sodergren, M.H. CBD Effects on TRPV1 Signaling Pathways in Cultured DRG Neurons. J. Pain Res. 2020, 13, 2269–2278. [Google Scholar] [CrossRef] [PubMed]
  48. Nazıroğlu, M.; Demirdaş, A. Psychiatric Disorders and TRP Channels: Focus on Psychotropic Drugs. Curr. Neuropharmacol. 2015, 13, 248–257. [Google Scholar] [CrossRef]
  49. Marsch, R.; Foeller, E.; Rammes, G.; Bunck, M.; Kössl, M.; Holsboer, F.; Zieglgänsberger, W.; Landgraf, R.; Lutz, B.; Wotjak, C.T. Reduced Anxiety, Conditioned Fear, and Hippocampal Long-Term Potentiation in Transient Receptor Potential Vanilloid Type 1 Receptor-Deficient Mice. J. Neurosci. 2007, 27, 832–839. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. Kasckow, J.W.; Mulchahey, J.J.; Geracioti, T.D. Effects of the Vanilloid Agonist Olvanil and Antagonist Capsazepine on Rat Behaviors. Prog. Neuro Psychopharmacol. Biol. Psychiatry 2004, 28, 291–295. [Google Scholar] [CrossRef]
  51. Santos, C.J.P.A.; Stern, C.A.J.; Bertoglio, L.J. Attenuation of Anxiety-Related Behaviour after the Antagonism of Transient Receptor Potential Vanilloid Type 1 Channels in the Rat Ventral Hippocampus. Behav. Pharmacol. 2008, 19, 357–360. [Google Scholar] [CrossRef]
  52. Aguiar, D.C.; Almeida-Santos, A.F.; Moreira, F.A.; Guimarães, F.S. Involvement of TRPV1 Channels in the Periaqueductal Grey on the Modulation of Innate Fear Responses. Acta Neuropsychiatr. 2015, 27, 97–105. [Google Scholar] [CrossRef] [PubMed]
  53. Mezey, É.; Tóth, Z.E.; Cortright, D.N.; Arzubi, M.K.; Krause, J.E.; Elde, R.; Guo, A.; Blumberg, P.M.; Szallasi, A. Distribution of MRNA for Vanilloid Receptor Subtype 1 (VR1), and VR1-like Immunoreactivity, in the Central Nervous System of the Rat and Human. Proc. Natl. Acad. Sci. USA 2000, 97, 3655–3660. [Google Scholar] [CrossRef]
  54. Jia, Y.-F.; Li, Y.-C.; Tang, Y.-P.; Cao, J.; Wang, L.-P.; Yang, Y.-X.; Xu, L.; Mao, R.-R. Interference of TRPV1 Function Altered the Susceptibility of PTZ-Induced Seizures. Front. Cell. Neurosci. 2015, 9, 20. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Sun, F.-J.; Guo, W.; Zheng, D.-H.; Zhang, C.-Q.; Li, S.; Liu, S.-Y.; Yin, Q.; Yang, H.; Shu, H.-F. Increased Expression of TRPV1 in the Cortex and Hippocampus from Patients with Mesial Temporal Lobe Epilepsy. J. Mol. Neurosci. 2013, 49, 182–193. [Google Scholar] [CrossRef] [PubMed]
  56. Alawi, K.M.; Russell, F.A.; Aubdool, A.A.; Srivastava, S.; Riffo-Vasquez, Y.; Baldissera, L.; Thakore, P.; Saleque, N.; Fernandes, E.S.; Walsh, D.A.; et al. Transient Receptor Potential Canonical 5 (TRPC5) Protects against Pain and Vascular Inflammation in Arthritis and Joint Inflammation. Ann. Rheum. Dis. 2017, 76, 252–260. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Lawhorn, B.G.; Brnardic, E.J.; Behm, D.J. Recent Advances in TRPV4 Agonists and Antagonists. Bioorg. Med. Chem. Lett. 2020, 30, 127022. [Google Scholar] [CrossRef]
  58. Sadler, K.E.; Moehring, F.; Shiers, S.I.; Laskowski, L.J.; Mikesell, A.R.; Plautz, Z.R.; Brezinski, A.N.; Mecca, C.M.; Dussor, G.; Price, T.J.; et al. Transient Receptor Potential Canonical 5 Mediates Inflammatory Mechanical and Spontaneous Pain in Mice. Sci. Transl. Med. 2021, 13, eabd7702. [Google Scholar] [CrossRef]
  59. Yang, L.; Liu, C.; Li, W.; Ma, Y.; Huo, S.; Ozathaley, A.; Ren, J.; Yuan, W.; Ni, H.; Li, D.; et al. Depression-like Behavior Associated with E/I Imbalance of MPFC and Amygdala without TRPC Channels in Mice of Knockout IL-10 from Microglia. Brain Behav. Immun. 2021, 97, 68–78. [Google Scholar] [CrossRef]
  60. Just, S.; Chenard, B.L.; Ceci, A.; Strassmaier, T.; Chong, J.A.; Blair, N.T.; Gallaschun, R.J.; Del Camino, D.; Cantin, S.; D’Amours, M.; et al. Treatment with HC-070, a Potent Inhibitor of TRPC4 and TRPC5, Leads to Anxiolytic and Antidepressant Effects in Mice. PLoS ONE 2018, 13, e0191225. [Google Scholar] [CrossRef]
  61. Nagib, M.M.; Zhang, S.; Yasmen, N.; Li, L.; Hou, R.; Yu, Y.; Boda, V.K.; Wu, Z.; Li, W.; Jiang, J. Inhibition of TRPC3 Channels by a Novel Pyrazole Compound Confers Antiseizure Effects. Epilepsia 2022, 63, 1003–1015. [Google Scholar] [CrossRef]
  62. Lee, K.; Jo, Y.Y.; Chung, G.; Jung, J.H.; Kim, Y.H.; Park, C.-K. Functional Importance of Transient Receptor Potential (TRP) Channels in Neurological Disorders. Front. Cell Dev. Biol. 2021, 9, 611773. [Google Scholar] [CrossRef]
  63. Zang, Z.; Li, S.; Zhang, W.; Chen, X.; Zheng, D.; Shu, H.; Guo, W.; Zhao, B.; Shen, K.; Wei, Y.; et al. Expression Patterns of TRPC1 in Cortical Lesions from Patients with Focal Cortical Dysplasia. J. Mol. Neurosci. 2015, 57, 265–272. [Google Scholar] [CrossRef] [PubMed]
  64. Yam, M.F.; Loh, Y.C.; Tan, C.S.; Khadijah Adam, S.; Abdul Manan, N.; Basir, R. General Pathways of Pain Sensation and the Major Neurotransmitters Involved in Pain Regulation. Int. J. Mol. Sci. 2018, 19, 2164. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Pitake, S.; Middleton, L.J.; Abdus-Saboor, I.; Mishra, S.K. Inflammation Induced Sensory Nerve Growth and Pain Hypersensitivity Requires the N-Type Calcium Channel Cav2.2. Front. Neurosci. 2019, 13, 1009. [Google Scholar] [CrossRef] [PubMed]
  66. Jimenez, I.; Prado, Y.; Marchant, F.; Otero, C.; Eltit, F.; Cabello-Verrugio, C.; Cerda, O.; Simon, F. TRPM Channels in Human Diseases. Cells 2020, 9, 2604. [Google Scholar] [CrossRef]
  67. Behrendt, M. Transient Receptor Potential Channels in the Context of Nociception and Pain—Recent Insights into TRPM3 Properties and Function. Biol. Chem. 2019, 400, 917–926. [Google Scholar] [CrossRef] [PubMed]
  68. Vriens, J.; Owsianik, G.; Hofmann, T.; Philipp, S.E.; Stab, J.; Chen, X.; Benoit, M.; Xue, F.; Janssens, A.; Kerselaers, S.; et al. TRPM3 Is a Nociceptor Channel Involved in the Detection of Noxious Heat. Neuron 2011, 70, 482–494. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Descoeur, J.; Pereira, V.; Pizzoccaro, A.; Francois, A.; Ling, B.; Maffre, V.; Couette, B.; Busserolles, J.; Courteix, C.; Noel, J.; et al. Oxaliplatin-Induced Cold Hypersensitivity Is Due to Remodelling of Ion Channel Expression in Nociceptors. EMBO Mol. Med. 2011, 3, 266–278. [Google Scholar] [CrossRef]
  70. Burgos-Vega, C.C.; Ahn, D.D.-U.; Bischoff, C.; Wang, W.; Horne, D.; Wang, J.; Gavva, N.; Dussor, G. Meningeal Transient Receptor Potential Channel M8 Activation Causes Cutaneous Facial and Hindpaw Allodynia in a Preclinical Rodent Model of Headache. Cephalalgia 2016, 36, 185–193. [Google Scholar] [CrossRef]
  71. Wu, L.; Zhang, J.; Zhou, F.; Zhang, P. Increased Transient Receptor Potential Melastatin 8 Expression in the Development of Bladder Pain in Patients with Interstitial Cystitis/Bladder Pain Syndrome. Urology 2020, 146, 301.e1–301.e6. [Google Scholar] [CrossRef]
  72. Mukerji, G.; Yiangou, Y.; Corcoran, S.L.; Selmer, I.S.; Smith, G.D.; Benham, C.D.; Bountra, C.; Agarwal, S.K.; Anand, P. Cool and Menthol Receptor TRPM8 in Human Urinary Bladder Disorders and Clinical Correlations. BMC Urol. 2006, 6, 6. [Google Scholar] [CrossRef] [Green Version]
  73. Nakao, A.; Matsunaga, Y.; Hayashida, K.; Takahashi, N. Role of Oxidative Stress and Ca2+ Signaling in Psychiatric Disorders. Front. Cell Dev. Biol. 2021, 9, 615569. [Google Scholar] [CrossRef]
  74. Rossetti, A.C.; Paladini, M.S.; Riva, M.A.; Molteni, R. Oxidation-Reduction Mechanisms in Psychiatric Disorders: A Novel Target for Pharmacological Intervention. Pharmacol. Ther. 2020, 210, 107520. [Google Scholar] [CrossRef] [PubMed]
  75. Fernandes, B.S.; Dean, O.M.; Dodd, S.; Malhi, G.S.; Berk, M. N-Acetylcysteine in Depressive Symptoms and Functionality: A Systematic Review and Meta-Analysis. J. Clin. Psychiatry 2016, 77, e457–e466. [Google Scholar] [CrossRef] [PubMed]
  76. Sepehrmanesh, Z.; Heidary, M.; Akasheh, N.; Akbari, H.; Heidary, M. Therapeutic Effect of Adjunctive N-Acetyl Cysteine (NAC) on Symptoms of Chronic Schizophrenia: A Double-Blind, Randomized Clinical Trial. Prog. Neuro Psychopharmacol. Biol. Psychiatry 2018, 82, 289–296. [Google Scholar] [CrossRef] [PubMed]
  77. Sakaguchi, R.; Mori, Y. Transient Receptor Potential (TRP) Channels: Biosensors for Redox Environmental Stimuli and Cellular Status. Free Radic. Biol. Med. 2020, 146, 36–44. [Google Scholar] [CrossRef] [PubMed]
  78. Olah, M.E.; Jackson, M.F.; Li, H.; Perez, Y.; Sun, H.-S.; Kiyonaka, S.; Mori, Y.; Tymianski, M.; MacDonald, J.F. Ca2+-Dependent Induction of TRPM2 Currents in Hippocampal Neurons. J. Physiol. 2009, 587, 965–979. [Google Scholar] [CrossRef] [PubMed]
  79. Xie, Y.-F.; Belrose, J.C.; Lei, G.; Tymianski, M.; Mori, Y.; Macdonald, J.F.; Jackson, M.F. Dependence of NMDA/GSK-3β Mediated Metaplasticity on TRPM2 Channels at Hippocampal CA3-CA1 Synapses. Mol. Brain 2011, 4, 44. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  80. Thippeswamy, H.; Davies, W. A New Molecular Risk Pathway for Postpartum Mood Disorders: Clues from Steroid Sulfatase-Deficient Individuals. Arch. Women’s Ment. Health 2021, 24, 391–401. [Google Scholar] [CrossRef]
  81. Maddaloni, G.; Migliarini, S.; Napolitano, F.; Giorgi, A.; Nazzi, S.; Biasci, D.; De Felice, A.; Gritti, M.; Cavaccini, A.; Galbusera, A.; et al. Serotonin Depletion Causes Valproate-Responsive Manic-like Condition and Increased Hippocampal Neuroplasticity That Are Reversed by Stress. Sci. Rep. 2018, 8, 11847. [Google Scholar] [CrossRef]
  82. Steinlein, O.K. Calcium Signaling and Epilepsy. Cell Tissue Res. 2014, 357, 385–393. [Google Scholar] [CrossRef]
  83. Tucholski, J.; Roth, K.A.; Johnson, G.V.W. Tissue Transglutaminase Overexpression in the Brain Potentiates Calcium-Induced Hippocampal Damage. J. Neurochem. 2006, 97, 582–594. [Google Scholar] [CrossRef] [PubMed]
  84. Arundine, M.; Tymianski, M. Molecular Mechanisms of Calcium-Dependent Neurodegeneration in Excitotoxicity. Cell Calcium 2003, 34, 325–337. [Google Scholar] [CrossRef]
  85. Muona, M.; Berkovic, S.F.; Dibbens, L.M.; Oliver, K.L.; Maljevic, S.; Bayly, M.A.; Joensuu, T.; Canafoglia, L.; Franceschetti, S.; Michelucci, R.; et al. A Recurrent de Novo Mutation in KCNC1 Causes Progressive Myoclonus Epilepsy. Nat. Genet. 2015, 47, 39–46. [Google Scholar] [CrossRef] [PubMed]
  86. Wengert, E.R.; Patel, M.K. The Role of the Persistent Sodium Current in Epilepsy. Epilepsy Curr. 2020, 21, 40–47. [Google Scholar] [CrossRef]
  87. Jeong, J.H.; Lee, S.H.; Kho, A.R.; Hong, D.K.; Kang, D.H.; Kang, B.S.; Park, M.K.; Choi, B.Y.; Choi, H.C.; Lim, M.-S.; et al. The Transient Receptor Potential Melastatin 7 (TRPM7) Inhibitors Suppress Seizure-Induced Neuron Death by Inhibiting Zinc Neurotoxicity. Int. J. Mol. Sci. 2020, 21, 7897. [Google Scholar] [CrossRef]
  88. Stawicki, T.M.; Zhou, K.; Yochem, J.; Chen, L.; Jin, Y. TRPM Channels Modulate Epileptic-like Convulsions via Systemic Ion Homeostasis. Curr. Biol. 2011, 21, 883–888. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Morgans, C.W.; Zhang, J.; Jeffrey, B.G.; Nelson, S.M.; Burke, N.S.; Duvoisin, R.M.; Brown, R.L. TRPM1 Is Required for the Depolarizing Light Response in Retinal ON-Bipolar Cells. Proc. Natl. Acad. Sci. USA 2009, 106, 19174–19178. [Google Scholar] [CrossRef] [Green Version]
  90. Jara-Oseguera, A.; Simon, S.A.; Rosenbaum, T. TRPV1: On the Road to Pain Relief. Curr. Mol. Pharmacol. 2008, 1, 255–269. [Google Scholar] [CrossRef]
  91. Pickering, G.; Creveaux, I.; Macian, N.; Pereira, B. Paracetamol and Pain Modulation by TRPV1, UGT2B15, SULT1A1 Genotypes: A Randomized Clinical Trial in Healthy Volunteers. Pain Med. 2020, 21, 661–669. [Google Scholar] [CrossRef]
  92. Marrone, M.C.; Morabito, A.; Giustizieri, M.; Chiurchiù, V.; Leuti, A.; Mattioli, M.; Marinelli, S.; Riganti, L.; Lombardi, M.; Murana, E.; et al. TRPV1 Channels Are Critical Brain Inflammation Detectors and Neuropathic Pain Biomarkers in Mice. Nat. Commun. 2017, 8, 15292. [Google Scholar] [CrossRef] [Green Version]
  93. Kim, Y.H.; Back, S.K.; Davies, A.J.; Jeong, H.; Jo, H.J.; Chung, G.; Na, H.S.; Bae, Y.C.; Kim, S.J.; Kim, J.S.; et al. TRPV1 in GABAergic Interneurons Mediates Neuropathic Mechanical Allodynia and Disinhibition of the Nociceptive Circuitry in the Spinal Cord. Neuron 2012, 74, 640–647. [Google Scholar] [CrossRef] [Green Version]
  94. Kim, Y.S.; Chu, Y.; Han, L.; Li, M.; Li, Z.; LaVinka, P.C.; Sun, S.; Tang, Z.; Park, K.; Caterina, M.J.; et al. Central Terminal Sensitization of TRPV1 by Descending Serotonergic Facilitation Modulates Chronic Pain. Neuron 2014, 81, 873–887. [Google Scholar] [CrossRef] [Green Version]
  95. Schilling, J.M.; Hughes, C.G.; Wallace, M.S.; Sexton, M.; Backonja, M.; Moeller-Bertram, T. Cannabidiol as a Treatment for Chronic Pain: A Survey of Patients’ Perspectives and Attitudes. J. Pain Res. 2021, 14, 1241–1250. [Google Scholar] [CrossRef]
  96. Mnich, K.; Finn, D.P.; Dowd, E.; Gorman, A.M. Inhibition by Anandamide of 6-Hydroxydopamine-Induced Cell Death in PC12 Cells. Int. J. Cell Biol. 2010, 2010, 818497. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Kumar, H.; Lee, S.-H.; Kim, K.-T.; Zeng, X.; Han, I. TRPV4: A Sensor for Homeostasis and Pathological Events in the CNS. Mol. Neurobiol. 2018, 55, 8695–8708. [Google Scholar] [CrossRef] [PubMed]
  98. Fu, S.; Meng, H.; Inamdar, S.; Das, B.; Gupta, H.; Wang, W.; Thompson, C.L.; Knight, M.M. Activation of TRPV4 by Mechanical, Osmotic or Pharmaceutical Stimulation Is Anti-Inflammatory Blocking IL-1β Mediated Articular Cartilage Matrix Destruction. Osteoarthr. Cartil. 2021, 29, 89–99. [Google Scholar] [CrossRef] [PubMed]
  99. Ding, X.-L.; Wang, Y.-H.; Ning, L.-P.; Zhang, Y.; Ge, H.-Y.; Jiang, H.; Wang, R.; Yue, S.-W. Involvement of TRPV4-NO-CGMP-PKG Pathways in the Development of Thermal Hyperalgesia Following Chronic Compression of the Dorsal Root Ganglion in Rats. Behav. Brain Res. 2010, 208, 194–201. [Google Scholar] [CrossRef] [PubMed]
  100. Lin, X.-Y.; Yang, J.; Li, H.-M.; Hu, S.-J.; Xing, J.-L. Dorsal Root Ganglion Compression as an Animal Model of Sciatica and Low Back Pain. Neurosci. Bull. 2012, 28, 618–630. [Google Scholar] [CrossRef] [Green Version]
  101. Brooks, C.A.; Barton, L.S.; Behm, D.J.; Brnardic, E.J.; Costell, M.H.; Holt, D.A.; Jolivette, L.J.; Matthews, J.M.; McAtee, J.J.; McCleland, B.W.; et al. Discovery of GSK3527497: A Candidate for the Inhibition of Transient Receptor Potential Vanilloid-4 (TRPV4). J. Med. Chem. 2019, 62, 9270–9280. [Google Scholar] [CrossRef] [PubMed]
  102. Tian, Y.-H.; Lee, S.-Y.; Kim, H.-C.; Jang, C.-G. Repeated Methamphetamine Treatment Increases Expression of TRPV1 MRNA in the Frontal Cortex but Not in the Striatum or Hippocampus of Mice. Neurosci. Lett. 2010, 472, 61–64. [Google Scholar] [CrossRef]
  103. Blednov, Y.A.; Harris, R.A. Deletion of Vanilloid Receptor (TRPV1) in Mice Alters Behavioral Effects of Ethanol. Neuropharmacology 2009, 56, 814–820. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Manna, S.S.S.; Umathe, S.N. Involvement of Transient Receptor Potential Vanilloid Type 1 Channels in the Pro-Convulsant Effect of Anandamide in Pentylenetetrazole-Induced Seizures. Epilepsy Res. 2012, 100, 113–124. [Google Scholar] [CrossRef] [PubMed]
  105. Nazıroğlu, M.; Övey, İ.S. Involvement of Apoptosis and Calcium Accumulation through TRPV1 Channels in Neurobiology of Epilepsy. Neuroscience 2015, 293, 55–66. [Google Scholar] [CrossRef] [PubMed]
  106. Qin, N.; Neeper, M.P.; Liu, Y.; Hutchinson, T.L.; Lubin, M.L.; Flores, C.M. TRPV2 Is Activated by Cannabidiol and Mediates CGRP Release in Cultured Rat Dorsal Root Ganglion Neurons. J. Neurosci. 2008, 28, 6231–6238. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Jones, N.A.; Hill, A.J.; Smith, I.; Bevan, S.A.; Williams, C.M.; Whalley, B.J.; Stephens, G.J. Cannabidiol Displays Antiepileptiform and Antiseizure Properties In Vitro and In Vivo. J. Pharmacol. Exp. Ther. 2010, 332, 569–577. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Liu, Y.; Liu, Y.; Limjunyawong, N.; Narang, C.; Jamaldeen, H.; Yu, S.; Patiram, S.; Nie, H.; Caterina, M.J.; Dong, X.; et al. Sensory Neuron–Expressed TRPC3 Mediates Acute and Chronic Itch. Pain 2023, 164, 98–110. [Google Scholar] [CrossRef] [PubMed]
  109. Alkhani, H.; Ase, A.R.; Grant, R.; O’Donnell, D.; Groschner, K.; Séguéla, P. Contribution of TRPC3 to Store-Operated Calcium Entry and Inflammatory Transductions in Primary Nociceptors. Mol. Pain 2014, 10, 43. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  110. Kress, M.; Karasek, J.; Ferrer-Montiel, A.V.; Scherbakov, N.; Haberberger, R.V. TRPC Channels and Diacylglycerol Dependent Calcium Signaling in Rat Sensory Neurons. Histochem. Cell Biol. 2008, 130, 655–667. [Google Scholar] [CrossRef]
  111. De Sousa Valente, J.; Alawi, K.M.; Keringer, P.; Bharde, S.; Ayaz, F.; Saleque, N.; Kodji, X.; Thapa, D.; Argunhan, F.; Brain, S.D. Examining the Role of Transient Receptor Potential Canonical 5 (TRPC5) in Osteoarthritis. Osteoarthr. Cartil. Open 2020, 2, 100119. [Google Scholar] [CrossRef]
  112. Riccio, A.; Li, Y.; Moon, J.; Kim, K.-S.; Smith, K.S.; Rudolph, U.; Gapon, S.; Yao, G.L.; Tsvetkov, E.; Rodig, S.J.; et al. Essential Role for TRPC5 in Amygdala Function and Fear-Related Behavior. Cell 2009, 137, 761–772. [Google Scholar] [CrossRef] [Green Version]
  113. Roedding, A.S.; Gao, A.F.; Au-Yeung, W.; Scarcelli, T.; Li, P.P.; Warsh, J.J. Effect of Oxidative Stress on TRPM2 and TRPC3 Channels in B Lymphoblast Cells in Bipolar Disorder. Bipolar Disord. 2012, 14, 151–161. [Google Scholar] [CrossRef] [PubMed]
  114. Andreopoulos, S.; Wasserman, M.; Woo, K.; Li, P.P.; Warsh, J.J. Chronic Lithium Treatment of B Lymphoblasts from Bipolar Disorder Patients Reduces Transient Receptor Potential Channel 3 Levels. Pharm. J. 2004, 4, 365–373. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Wang, D.; Yu, H.; Xu, B.; Xu, H.; Zhang, Z.; Ren, X.; Yuan, J.; Liu, J.; Guo, Y.; Spencer, P.S.; et al. TRPC1 Deletion Causes Striatal Neuronal Cell Apoptosis and Proteomic Alterations in Mice. Front. Aging Neurosci. 2018, 10, 72. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Yang, L.-P.; Jiang, F.-J.; Wu, G.-S.; Deng, K.; Wen, M.; Zhou, X.; Hong, X.; Zhu, M.X.; Luo, H.-R. Acute Treatment with a Novel TRPC4/C5 Channel Inhibitor Produces Antidepressant and Anxiolytic-Like Effects in Mice. PLoS ONE 2015, 10, e0136255. [Google Scholar] [CrossRef] [PubMed]
  117. Zhou, F.-W.; Roper, S.N. TRPC3 Mediates Hyperexcitability and Epileptiform Activity in Immature Cortex and Experimental Cortical Dysplasia. J. Neurophysiol. 2014, 111, 1227–1237. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Yu, Y.; Li, W.; Jiang, J. TRPC Channels as Emerging Targets for Seizure Disorders. Trends Pharmacol. Sci. 2022, 43, 787–798. [Google Scholar] [CrossRef]
  119. Sun, D.; Ma, H.; Ma, J.; Wang, J.; Deng, X.; Hu, C.; Deng, X. Canonical Transient Receptor Potential Channel 3 Contributes to Febrile Seizure Inducing Neuronal Cell Death and Neuroinflammation. Cell. Mol. Neurobiol. 2018, 38, 1215–1226. [Google Scholar] [CrossRef]
  120. Kim, J.-E.; Park, H.; Choi, S.-H.; Kong, M.-J.; Kang, T.-C. TRPC6-Mediated ERK1/2 Activation Increases Dentate Granule Cell Resistance to Status Epilepticus via Regulating Lon Protease-1 Expression and Mitochondrial Dynamics. Cells 2019, 8, 1376. [Google Scholar] [CrossRef] [Green Version]
  121. Jardín, I.; López, J.J.; Diez, R.; Sánchez-Collado, J.; Cantonero, C.; Albarrán, L.; Woodard, G.E.; Redondo, P.C.; Salido, G.M.; Smani, T.; et al. TRPs in Pain Sensation. Front. Physiol. 2017, 8, 392. [Google Scholar] [CrossRef] [Green Version]
  122. Julius, D. TRP Channels and Pain. Annu. Rev. Cell Dev. Biol. 2013, 29, 355–384. [Google Scholar] [CrossRef] [Green Version]
  123. Vowles, K.E.; McEntee, M.L.; Julnes, P.S.; Frohe, T.; Ney, J.P.; van der Goes, D.N. Rates of Opioid Misuse, Abuse, and Addiction in Chronic Pain: A Systematic Review and Data Synthesis. Pain 2015, 156, 569–576. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Krebs, E.E.; Gravely, A.; Nugent, S.; Jensen, A.C.; DeRonne, B.; Goldsmith, E.S.; Kroenke, K.; Bair, M.J.; Noorbaloochi, S. Effect of Opioid vs Nonopioid Medications on Pain-Related Function in Patients with Chronic Back Pain or Hip or Knee Osteoarthritis Pain: The SPACE Randomized Clinical Trial. JAMA 2018, 319, 872–882. [Google Scholar] [CrossRef] [PubMed]
  125. Izquierdo, C.; Martín-Martínez, M.; Gómez-Monterrey, I.; González-Muñiz, R. TRPM8 Channels: Advances in Structural Studies and Pharmacological Modulation. Int. J. Mol. Sci. 2021, 22, 8502. [Google Scholar] [CrossRef]
  126. Rosenberger, D.C.; Binzen, U.; Treede, R.-D.; Greffrath, W. The Capsaicin Receptor TRPV1 Is the First Line Defense Protecting from Acute Non Damaging Heat: A Translational Approach. J. Transl. Med. 2020, 18, 28. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Moreau, C.; Kirchberger, T.; Swarbrick, J.M.; Bartlett, S.J.; Fliegert, R.; Yorgan, T.; Bauche, A.; Harneit, A.; Guse, A.H.; Potter, B.V.L. Structure-Activity Relationship of Adenosine 5′-Diphosphoribose at the Transient Receptor Potential Melastatin 2 (TRPM2) Channel: Rational Design of Antagonists. J. Med. Chem. 2013, 56, 10079–10102. [Google Scholar] [CrossRef]
  128. Hayase, T. Differential Effects of TRPV1 Receptor Ligands against Nicotine-Induced Depression-like Behaviors. BMC Pharmacol. 2011, 11, 6. [Google Scholar] [CrossRef] [Green Version]
  129. Ho, K.W.; Ward, N.J.; Calkins, D.J. TRPV1: A Stress Response Protein in the Central Nervous System. Am. J. Neurodegener. Dis. 2012, 1, 1–14. [Google Scholar]
  130. Barker-Haliski, M.; White, H.S. Glutamatergic Mechanisms Associated with Seizures and Epilepsy. Cold Spring Harb. Perspect. Med. 2015, 5, a022863. [Google Scholar] [CrossRef] [Green Version]
  131. Krügel, U.; Straub, I.; Beckmann, H.; Schaefer, M. Primidone Inhibits TRPM3 and Attenuates Thermal Nociception in Vivo. Pain 2017, 158, 856–867. [Google Scholar] [CrossRef] [Green Version]
  132. Menigoz, A.; Boudes, M.; Yamada, K.; Cirrito, J.R.; Stewart, F.R.; Jiang, H.; Finn, M.B.; Holmes, B.B.; Binder, L.I.; Mandelkow, E.-M.; et al. The Expression Pattern of TRPV1 in Brain. J. Neurosci. 2011, 31, 13025–13027. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Visualization of the generic structure of the transient receptor potential (TRP) channel, consistent for most families within the class. The pore between the fifth and sixth transmembrane domains yields nonselective calcium ion permeability. TRPM4 and TRPM5 are two exceptions, as they are calcium impermeable channels (Created with BioRender.com).
Figure 1. Visualization of the generic structure of the transient receptor potential (TRP) channel, consistent for most families within the class. The pore between the fifth and sixth transmembrane domains yields nonselective calcium ion permeability. TRPM4 and TRPM5 are two exceptions, as they are calcium impermeable channels (Created with BioRender.com).
Ijms 24 04714 g001
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Yang, F.; Sivils, A.; Cegielski, V.; Singh, S.; Chu, X.-P. Transient Receptor Potential (TRP) Channels in Pain, Neuropsychiatric Disorders, and Epilepsy. Int. J. Mol. Sci. 2023, 24, 4714. https://doi.org/10.3390/ijms24054714

AMA Style

Yang F, Sivils A, Cegielski V, Singh S, Chu X-P. Transient Receptor Potential (TRP) Channels in Pain, Neuropsychiatric Disorders, and Epilepsy. International Journal of Molecular Sciences. 2023; 24(5):4714. https://doi.org/10.3390/ijms24054714

Chicago/Turabian Style

Yang, Felix, Andy Sivils, Victoria Cegielski, Som Singh, and Xiang-Ping Chu. 2023. "Transient Receptor Potential (TRP) Channels in Pain, Neuropsychiatric Disorders, and Epilepsy" International Journal of Molecular Sciences 24, no. 5: 4714. https://doi.org/10.3390/ijms24054714

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

Article Metrics

Back to TopTop