Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2024 Apr 1.
Published in final edited form as: Pharmacol Ther. 2023 Mar 16;244:108384. doi: 10.1016/j.pharmthera.2023.108384

Pharmacologically targeting transient receptor potential channels for seizures and epilepsy: Emerging preclinical evidence of druggability

Jianxiong Jiang 1,*, Ying Yu 1,*
PMCID: PMC10124570  NIHMSID: NIHMS1886063  PMID: 36933703

Abstract

As one of the most prevalent and disabling brain disorders, epilepsy is characterized by spontaneous seizures that result from aberrant, excessive hyperactivity of a group of highly synchronized brain neurons. Remarkable progress in epilepsy research and treatment over the first two decades of this century led to a dramatical expansion in the third-generation antiseizure drugs (ASDs). However, there are still over 30% of patients suffering from seizures resistant to the current medications, and the broad unbearable adversative effects of ASDs significantly impair the quality of life in about 40% of individuals affected by the disease. Prevention of epilepsy in those who are at high risks is another major unmet medical need, given that up to 40% of epilepsy patients are believed to have acquired causes. Therefore, it is important to identify novel drug targets that can facilitate the discovery and development of new therapies engaging unprecedented mechanisms of action that might overcome these significant limitations. Also over the last two decades, calcium signaling has been increasingly recognized as a key contributory factor in epileptogenesis of many aspects. The intracellular calcium homeostasis involves a variety of calcium-permeable cation channels, the most important of which perhaps are the transient receptor potential (TRP) ion channels. This review focuses on recent exciting advances in understanding of TRP channels in preclinical models of seizure disorders. We also provide emerging insights into the molecular and cellular mechanisms of TRP channels-engaged epileptogenesis that might lead to new antiseizure therapies, epilepsy prevention and modification, and even a cure.

Keywords: Antiseizure drug (ASD), Calcium signaling, Cation channels, Epileptogenesis, Excitotoxicity, Hyperexcitability, Seizures, Transient receptor potential (TRP) channels

1. Introduction

Afflicting about 65 million people globally, epilepsy is a group of chronic debilitating brain conditions that have diverse etiologies and can result from genetic mutations, acute brain injuries, autoimmune conditions, brain infections, and metabolic disorders. For a large proportion of patients, however, the causes of epilepsy are unknown. Regardless of diverse etiologies, epilepsy is commonly featured by spontaneous recurrent seizures due to aberrant, excessive hyperactivity of a group of highly synchronized brain neurons (Devinsky, et al., 2018). Seizures are abrupt, uncontrollable behavioral changes with a large range of signs, such as confusion, unusual feelings and sensations, loss of consciousness, stiffness, jerking, falling, etc. The onset of seizures can be generalized or focal, depending on whether the irregular neuronal activities begin in an extensive distribution affecting both hemispheres or in a more focalized brain areas involving only one hemisphere. In addition, epilepsy is commonly associated with many neurological comorbidities, such as anxiety and depression, cognitive deficits, and psychiatric disturbances, which along with seizures themselves severely impact the quality of life for patients (Devinsky, et al., 2018; Fisher, et al., 2017a; Fisher, et al., 2017b).

Despite the remarkable advances in epilepsy research and treatment in the first two decades of this century, leading to an expanding list of the third-generation antiseizure drugs (ASDs) (Carvill, Dulla, Lowenstein, & Brooks-Kayal, 2020; Loscher & Klein, 2020; Rho & White, 2018), there are still approximately 1/3 of epilepsy patients who are inadequately treated and suffer from pharmacoresistant seizures (Janmohamed, Brodie, & Kwan, 2020). Up to date, there is no undisputed evidence showing that the use of newer ASDs leads to more seizure-free patients than the established antiseizure medications (Chen, Brodie, Liew, & Kwan, 2018; Hauser, 2018). The chronic use of current ASDs can impose a wide range of adverse effects, most commonly, ataxia, attention disturbance, blurred vision, cognitive decline, confusion, constipation, depression, diarrhea, diplopia, dizziness, drowsiness, fatigue, headache, incoordination, nausea, sedation, tremor, vertigo, vomiting, and weight gain. These systemic and neurological side effects of ASDs further complicate seizure management, particularly when polytherapy is needed (Devinsky, et al., 2018). Moreover, the current ASDs, regardless of newer or established drugs, only interrupt seizures in patients who have been diagnosed with epilepsy, and up to date, they have not been demonstrated to prevent epilepsy after precipitating events or modify the disease progression (Galanopoulou, et al., 2021; Loscher, 2020).

Most ASDs suppress seizures through at least one of the four currently known mechanisms of action: 1) modulation of voltage-gated ion channels; 2) modulation of neurotransmitter release; 3) impediment of glutamatergic excitation; 4) augmentation of GABAergic inhibition (Loscher & Klein, 2020; Sills & Rogawski, 2020). For instance, ganaxolone, a recently approved ASD, is thought to control seizures in patients with the cyclin-dependent kinase-like 5 deficiency disorder via positively modulating the GABAA receptor (Yasmen, Sluter, Yu, & Jiang, 2023). However, these four classical antiseizure mechanisms may also underlie the adverse effects of ASDs (Sills & Rogawski, 2020), and the pharmacoresistant epilepsy is likely developed due to the target-site alterations (Janmohamed, et al., 2020). As such, identification of novel drug targets may facilitate the development of safer antiseizure therapies that engage unprecedented mechanisms of drug action to terminate acute seizures in a more efficient way and may even prevent the development of epilepsy after precipitating brain insults or modify the progression of epileptic seizures (Dey, Kang, Qiu, Du, & Jiang, 2016; Varvel, Jiang, & Dingledine, 2015). Such efforts presumably will help to overcome the seemingly insurmountable limitations in the current epilepsy medications.

2. TRP cation channels

The transient receptor potential (TRP) ion channels are a group of tetrameric nonselective cation channels that are essential to the intracellular calcium homeostasis and contributes to many different physiological processes (Prakriya & Lewis, 2015). In mammals, this superfamily has 28 currently known members that can be classified into six subfamilies based on amino acid homologies: TRPA (ankyrin, A1), TRPC (canonical, C1-C7), TRPM (melastatin, M1-M8), TRPML (mucolipin, ML1-ML3), TRPP (polycystin, P1-P3), and TRPV (vanilloid, V1-V6) (Nilius & Szallasi, 2014; Samanta, Hughes, & Moiseenkova-Bell, 2018). However, in humans, TRPC2 is a pseudogene that is not expressed. Most members of TRPC, TRPM, and TRPV subfamilies are highly expressed within the brain (Wu, Sweet, & Clapham, 2010), where they can regulate a variety of cellular functions of both neurons and glial cells (Harteneck & Leuner, 2014). Particularly, the activation of TRP channels expressed by brain neurons can cause the depolarization of membrane potential to trigger the activation or inactivation of many different voltage-gated ion channels and modulate calcium signaling to regulate diverse cellular functions. In addition to these immediate actions, elevated calcium in neurons may cause some long-term effect through calcium-dependent gene regulation, which contributes to several forms of synaptic plasticity, leading to long-term alterations in the biological and pathophysiological behavior of neural circuits (Meza, Ancaten-Gonzalez, Chiu, & Chavez, 2022; Nilius & Szallasi, 2014). As such, mounting evidence from recent studies suggests that dysregulation of the TRP channel functions is involved in pathological events of various neurological and psychiatric diseases, including disorders caused by defects in genes encoding these cation channels, which are thus called hereditary TRP channelopathies (Koivisto, Belvisi, Gaudet, & Szallasi, 2022).

Deregulated calcium signaling is now well recognized as a major contributor to epileptic seizures largely because of its involvement in the aberrant synchronization of brain neurons that is thought to underlie the epileptogenic mechanisms (Penn, Segal, & Moses, 2016). Moreover, astrocytic calcium signals have been implicated in the facilitation of spreading the epileptic activities (Heuser, et al., 2018; Sasaki, et al., 2014). Particularly, calcium signaling can lead to profound effects on membrane excitability directly by calcium influx or through indirect mechanisms involving G protein-dependent pathways (Brini, Cali, Ottolini, & Carafoli, 2014; Kawamoto, Vivar, & Camandola, 2012; Yu, Nguyen, & Jiang, 2019). Among a variety of calcium-permeable cation channels that are involved in the intracellular calcium homeostasis and have been implicated in acute seizures and chronic epilepsy are TRP ion channels (Garcia-Rodriguez, Bravo-Tobar, Duarte, Barrio, & Saez, 2022). The rest of this review focuses on several TRP superfamily members (TRPV, TRPC, and TRPM) that have recently been extensively studied in various animal models for their pathogenic roles in epileptic seizures. The emerging pharmacological evidence is also highlighted to demonstrate their druggability for the next-generation therapies to stop acute seizures and potentially epileptogenesis.

3. TRPV channels

3.1. TRPV1

Involved in many important pathological and physiological processes, TRPV1 has long been implicated in a variety of conditions, such as ischemia, epilepsy, schizophrenia, psoriasis, pneumonia, and diabetes (Gladkikh, Sintsova, Leychenko, & Kozlov, 2021). For instance, TRPV1 in cortical and hippocampal tissues from patients with mesial temporal lobe epilepsy (MTLE) is upregulated at both mRNA and protein levels when compared to the normal control subjects and is mainly found in the somas and dendrites of neurons (NeuN+) and in astrocytes (GFAP+) but not in microglia (HLA-DR+) (Sun, et al., 2013). These findings together with its expression in both glutamatergic and GABAergic neurons suggest that TRPV1 might contribute to the neuropathogenesis of human MTLE, inspiring studies on TRPV1 as a potential therapeutic target in various preclinical models (Table 1). Interestingly, treatment with selective TRPV1 agonist OLDA dose-dependently facilitated the overall incidence of seizures induced by chemoconvulsant pentylenetetrazole (PTZ) in Wistar rats. Conversely, high-stage seizures and afterdischarge duration in rat amygdala kindling model were significantly reduced after the administration of TRPV1 antagonist AMG-9810 (Shirazi, et al., 2014). Likewise, systemic administration of TRPV1 agonist capsaicin in mice directly triggered tonic-clonic seizures. In contrast, intraperitoneal administration of TRPV1 antagonist capsazepine or genetic ablation of TRPV1 led to delayed latency to tonic-clonic seizures and reduced overall mortality in mice treated with PTZ for seizure induction (Jia, et al., 2015). These pharmacological studies suggest that the selective inhibition of TRPV1 channels might represent a novel therapeutic strategy for epilepsy (Table 1). In line with these findings, TRPV1 activation by capsaicin increased, whereas blocking TRPV1 with capsazepine reduced, the paroxysmal discharge following the maximal dentate activation (MDA) in Wistar rats. Moreover, capsaicin, when co-administered with WIN 55,212–2, an exogenous cannabinoid agonist for the CB1 receptor, was able to reduce its anticonvulsant effects. In contrast, co-administration of capsazepine potentiated the suppression by WIN 55,212–2 on the MDA response, supporting a possible crosstalk between TRPV1 and cannabinoid signaling in hyperexcitability-associated brain diseases (Carletti, Gambino, Rizzo, Ferraro, & Sardo, 2016).

Table 1.

Pharmacological studies on TRPV channels in animal models of seizures and epilepsy.

Pharmacological Targets Animal Models Therapeutic Agents Major Outcomes References
TRPV1 PTZ-induced seizures in adult male Wistar rats OLDA (0.01, 0.1 and 1 μg/animal, i.c.v.) Facilitated the overall incidence of seizures Shirazi et al., 2014
TRPV1 Amygdala kindling in adult male Wistar rats AMG-9810 (0.3, 3 and 30 μg/animal, i.c.v.) Reduced high-stage seizures and afterdischarge duration Shirazi et al., 2014
TRPV1 Adult male Kunming mice Capsaicin (40 mg/kg, i p.) Promoted tonic-clonic seizures Jia et al., 2015
TRPV1 PTZ-induced seizures in adult male Kunming mice Capsazepine (0.05 and 0.5 mg/kg, i.p.) Delayed the latency to tonic-clonic seizures and reduced mortality Jia et al., 2015
TRPV1 MDA in adult male Wistar rats Capsaicin (1, 5, and 10 mg/kg, i.p.) Increased paroxysmal discharge and MDA response Carletti et al., 2016
TRPV1 MDA in adult male Wistar rats Capsazepine (0.5, 1, and 2 mg/kg, i.p.) Reduced paroxysmal discharge and MDA response Carletti et al., 2016
TRPV1 PTZ-induced seizures in adult male ICR mice Capsazepine (1, and 3 mg/kg, i.p.) Showed antiseizure activity alone but reduced anticonvulsant effects of acetaminophen Suemaru et al., 2018
TRPV1 PTZ-induced seizures in adult male ICR mice AMG-9810 (1, 3, and 10 mg/kg, i.p.) Showed antiseizure activity alone but reduced anticonvulsant effects of acetaminophen Suemaru et al., 2018
TRPV1 Lithium/pilocarpin e-induced seizures in immature male SD rats Capsaicin (0.1 mg/kg, i.p.) Increased neuronal death in hippocampus and reversed the neuroprotection of dexmedetomidi ne after SE Tan et al., 2020
TRPV1 HT-induced seizures in larval zebrafish Capsazepine (100 μM) Had no effect on HT seizure activity Hunt et al., 2012
TRPV4 HT-induced seizures in larval zebrafish RN-1734 (0.5–1 mM) Produced a significant reduction of HT seizure activity Hunt et al., 2012
TRPV4 Pilocarpine-induced seizures in young adult male ICR mice HC-067047 (10 μM/mouse, i.c.v.) Improved neuronal survival and diminished gliosis, inflammasome activation, and cytokine induction Wang et al., 2019
TRPV4 Pilocarpine-induced seizures in young adult male ICR mice HC-067047 (10 μM/mouse, i.c.v.) Delayed SE onset and reduced connexin expression Men et al., 2019
TRPV4 Pilocarpine-induced seizures in young male ICR mice HC-067047 (10 μM/mouse, i.c.v.) Attenuated the increased protein levels of Kv4.2 and KCHIP2 Xu et al., 2022
TRPV4 Pilocarpine-induced seizures in young male ICR mice GSK1016790 A (1 μM/mouse, i.c.v.) Increased hippocampal protein levels of Kv4.2 and KCHIP2 Xu et al., 2022
TRPV4 4-AP-induced seizures in young male adult C57BL/6 mice HC-067047 (10 μM/mouse, i.c.v.) Reduced seizure susceptibility, neuronal death, inflammatory cytokines, and improved neurobehavioral functions Zeng et al., 2022
TRPV4 4-AP-induced seizures in young male adult C57BL/6 mice GSK1016790 A (1 μM/mouse, i.c.v.) Increased animal mortality Zeng et al., 2022

Abbreviations: HT, hyperthermia; ICR, Institute of Cancer Research; i.c.v., intracerebroventricular; i.p., intraperitoneal; KCHIP, K+ channel interacting protein; MDA, maximal dentate activation; PTZ, pentylenetetrazole; SD, Sprague-Dawley; SE, status epilepticus.

In contrast to its well-studied proconvulsive roles discussed above, TRPV1 activation has also been reported to be associated with anticonvulsant effects in several preclinical models. For example, global congenital deletion of TRPV1 decreased the vulnerability to PTZ-induced seizures after repetitive hyperthermia challenges in neonatal mice (Kong, et al., 2014), suggestive of a role of TRPV1 in the pathogenesis of febrile seizures observed in newborns. However, loss of TRPV1 function prevented the excessive microglial activation and migration triggered by hyperthermic stimulation. It was further found that stimulation of microglial TRPV1 enhanced seizure susceptibility by decreasing the anti-inflammatory and neuroprotective effects of transforming growth factor β1 in microglia via interacting with toll-like receptor 4 (Kong, et al., 2019). Interestingly, systemic treatment with TRPV1 antagonists capsazepine or AMG-9810 alone reduced PTZ-provoked seizures in mice in a dose-dependent manner but decreased the anticonvulsant effects of acetaminophen in the same animal model (Suemaru, Yoshikawa, Aso, & Watanabe, 2018), suggesting that TRPV1 might be responsible for the anticonvulsant action of acetaminophen. In a lithium/pilocarpine model of status epilepticus (SE), treatment with capsaicin at a relatively low dose exacerbated neuronal death in the hippocampus and reversed neuroprotective effects of dexmedetomidine in immature rats (Tan, et al., 2020), indicating that TRPV1 activation might contribute to SE-induced neuronal death in juvenile animals. Taken together, whether TRPV1 activation leads to beneficial or detrimental effects appears to be determined by the seizure models, animal ages, and particularly, its cell type-specific distributions.

TRPV1 in the brain was induced in mice that experienced recurrent febrile seizures, particularly in activated microglia (Iba1+) (Kong, et al., 2019). This result seemingly contradicts early finding that TRPV1 was not expressed in HLA-DR+ microglia in temporal cortical tissues from MTLE patients (Sun, et al., 2013), but might be explained by the differences in seizure etiology (MTLE vs. febrile seizures), species (human vs. mouse), and microglial biomarkers (HLA-DR vs. Iba1) used in these two studies. Particularly, HLA-DR highlights the immune activation and response to brain tissue damage, whereas Iba1 emphasizes more on the cellular structure (Hendrickx, van Eden, Schuurman, Hamann, & Huitinga, 2017). Considering its discrete expression in neurons, astrocytes, and microglia in the epileptic brain (Kong, et al., 2019; Sun, et al., 2013), future studies engaging cell-type specific gene deletion (neuronal vs. microglial) are required to fully understand the complex roles of TRPV1 (anticonvulsive vs. proconvulsive) in seizure generation and subsequent neuropathogenesis.

3.2. TRPV4

Initially identified as a nonselective cationic channel permeable to calcium, TRPV4 is now known as a polymodal ionotropic receptor with a nonselective cation channel that is also permeable to potassium, magnesium, and sodium, and can be activated by a broad variety of stimuli (Zeng, Kong, Chen, & Peng, 2023). Its wide-ranging biological roles are associated with many physiological and pathological processes in the central nervous system (CNS). TRPV4 has long been implicated in the development of epilepsy of various origins, where it shows abnormal expression and activities. Both TRPV1 and TRPV4 were found upregulated in larval zebrafish after hyperthermia-induced seizures, but only the TRPV4 inhibition by RN1734, not the TRPV1 blockade by capsazepine, produced a significant reduction in hyperthermic seizure activities, an outcome that was recapitulated by NMDA receptor antagonists MK-801 and ifenprodil (Hunt, Hortopan, Gillespie, & Baraban, 2012). TRPV4 was also found elevated in the hippocampus after pilocarpine-induced SE in mice (Men, et al., 2019). Intracerebroventricular injection of TRPV4 agonist GSK1016790A in mice increased NLRP3 inflammasome activation, gliosis featured by upregulated Iba1 and GFPA, and pro-inflammatory cytokines IL-1β, IL-6 and TNF-α, and neuronal death in the hippocampus. On the contrary, intracerebroventricular injection of TRPV4 antagonist HC-067047 markedly improved neuronal survival after pilocarpine-induced SE and abolished SE-provoked gliosis, NLRP3 inflammasome activation, and cytokine induction (Wang, et al., 2019). Given that the pharmacological inhibition of TRPV4 by HC-067047 also delayed the latency to SE, suggestive of an anticonvulsant effect (Men, et al., 2019), the broad anti-inflammatory and neuroprotective effects from HC-067047 treatment in mouse SE model are likely secondary to its antiseizure action (Table 1). Interestingly, TRPV4 activation by GSK1016790A has been shown to enhance the expression of voltage-gated potassium channel Kv4.2 and potassium channel interacting protein (KCHIP) in the hippocampus following pilocarpine-induced SE in mice, which was attenuated by TRPV4 inhibitor HC-067047 (Table 1). The TRPV4 activity-mediated elevation of Kv4.2 and KCHIP is thought to lead to an increase in the rapidly inactivating potassium current in hippocampal pyramidal neurons, which likely contributes to hyperexcitability during the early stage of epileptogenesis (Xu, et al., 2022).

Neuronal excitation can evoke astrocytic calcium transients, which in turn can modulate neuronal excitability and regulate synaptic transmission. As such, TRPV4 expression was found in approximately 30% of astrocytes within the brain, and the activation of TRPV4-expressing astrocytes may lead to the release of glutamate, which can increase excitatory gliotransmission to enhance synaptic transmission and neuronal excitability (Shibasaki, Ikenaka, Tamalu, Tominaga, & Ishizaki, 2014). Both TRPV4 and GFAP were upregulated in the hippocampus following 4-AP-induced seizures in mice, and the substantial colocalization between these two proteins suggests that astrocytic activation could be TRPV4 activation-dependent. TRPV4 inhibition by HC-067047 reduced susceptibility of mice to 4-AP-induced seizures and several key pro-inflammatory mediators, whereas treatment with GSK1016790A aggravated animal mortality (Table 1) (Zeng, et al., 2022). These findings suggest that targeting astrocytic TRPV4 activation might provide a therapeutic strategy to treat acute seizures and development of epilepsy. In addition to the well-recognized neuroinflammatory mechanisms involving astrocytes, TRPV4 channels might regulate neuronal excitability via its activation by brain temperature. The brain temperature in epileptogenic foci was found to be dramatically higher than that in other normal regions, and this increase in temperature was critical for the progression of disease where TRPV4 played a fundamental role (Shibasaki, et al., 2020). Thus, it is highly likely that cooling treatment at the epileptogenic zones might be able to suppress epileptic discharges via inhibiting TRPV4 channels.

4. TRPC channels

4.1. TRPC3

TRPC3 is the second identified TRPC subfamily member and shares about 74% protein sequence identity with TRPC6 and TRPC7 (Yu, Li, & Jiang, 2022). TRPC3-containing cation channels are abundantly found in the brain, where they are mostly detected in the hippocampus, cortex, cerebellum, and substantia nigra and play important physiological functions (Hartmann, et al., 2008; Li, Calfa, Inoue, Amaral, & Pozzo-Miller, 2010; Mitsumura, Hosoi, Furuya, & Hirai, 2011; Roedding, et al., 2009; Zhou & Lee, 2011). However, recent evidence suggests that TRPC3 activity also contributes to many neuropathogenic processes, particularly in the epileptic brain. Congenital deletion of TRPC3 led to a reduction in the severity and duration of convulsions in mice treated by pilocarpine and markedly decreased the overall electroencephalogram (EEG) power and theta wave activities of chemoconvulsant SE (Phelan, et al., 2017). These findings demonstrated that TRPC3 activation might be essential to the seizure generation and motivated follow-on pharmacological studies to determine the feasibility of targeting TRPC3 to interrupt acute seizures and epileptogenesis.

Pyr3, a pyrazole compound that was identified as a highly potent and selective TRPC3 inhibitor (Kiyonaka, et al., 2009), when systemically administered in mice, led to a marked reduction in the overall root-mean-square power of SE induced by pilocarpine (Table 2). EEG recording revealed that TRPC3 inhibition by Pyr3 also specifically attenuated the theta activities during the entire course of SE (Phelan, et al., 2017). Interestingly, among all TRPC subtypes, only TRPC3 was significantly elevated in the hippocampus by febrile seizures induced in rats. Moreover, direct microinjection of Pyr3 into the hippocampus reduced the seizure severity and duration and largely prevented the febrile seizure-associated brain cell death and neuroinflammation (Sun, et al., 2018). Compound JW-65 was developed as a chemical analog of Pyr3 with higher metabolic stability and less toxicity (Zhang, et al., 2021), allowing high systemic doses in animal models, which would not be possible for Pyr3. Indeed, intraperitoneal administration of JW-65 strikingly diminished the behavioral seizures in pilocarpine-treated mice, and the results were validated by EEG recording and dose-dependently reproduced in a mouse PTZ model (Table 2). Remarkably, the antiseizure effects by JW-65 appeared comparable to that caused by the genetic deletion of TRPC3 in the same animal model (Nagib, et al., 2022; Phelan, et al., 2017), validating the TRPC3 as a pharmacological target for new treatment of seizures.

Table 2.

Pharmacological studies on TRPC channels in models of seizures and epilepsy.

Pharmacological Targets Animal Models Therapeutic Agents Major Outcomes References
TRPC3 Pilocarpine-induced seizures in young adult male SD rats Pyr3 (0.1 mg/kg/day, i.c.v.) Reduced hippocampal TRPC3 and protected hippocampal pyramidal neurons (Kim et al, 2013)
TRPC3 Pilocarpine-induced seizures in young adult male C57BL/6 mice Anti-TRPC3 antibody (1:6.5, 2 μL, i.c.v.) Decreased mossy fiber sprouting after SE (Zeng et al., 2015)
TRPC3 Pilocarpine-induced seizures in adult male C57BL/6 mice Pyr3 (3 mg/kg, i.p.) Reduced the overall RMS power of SE and theta activities during SE (Phelan et al., 2017)
TRPC3 HT-induced FS in juvenile SD rats Pyr3 (1 mg/kg, i.h.) Reduced neuronal cell death and neuroinflammation (Sun et al., 2018)
TRPC3 Pilocarpine-induced seizures in young adult male C57BL/6 mice JW-65 (100 mg/kg, i.p.) Diminished behavioral and electrographic seizures (Nagib et al., 2022)
TRPC3 PTZ-induced seizures in young adult male C57BL/6 mice JW-65 (20 or 100 mg/kg, i.p.) Elongated latencies to MJ and GTCS in a dose-dependent manner (Nagib et al., 2022)
TRPC5 Kainate-induced seizures in adult male SD rats NU6027 (100 μg/kg, i.p.) Had no effect on seizure severity but protected neurons in the piriform cortex and hippocampus (Park et al., 2019)
TRPC5 Post-TBI and PTZ-induced seizures in adult male C57BL/6J mice M084 (10 mg/kg, s.c.) Decreased post-TBI neuronal hyperexcitability and PTZ-provoked kindling (Carver et al., 2021)
TRPC6 Pilocarpine-induced seizures in young adult male SD rats Hyperforin (6 μM, 0.5 μL/h, i.c.v.) Increased hippocampal TRPC6 and protected hippocampal pyramidal neurons (Kim et al, 2013)
TRPC6 Pilocarpine-induced seizures in young adult male C57BL/6 mice Anti-TRPC6 antibody (2 μL, i.c.v.) Decreased the dendritic arborization and spine density of hippocampal CA3 neurons (Zeng et al., 2015)

Abbreviations: FS, febrile seizure; GTCS, generalized tonic-clonic seizure; HT, hyperthermia; i.c.v., intracerebroventricular; i.h., intrahippocampal; i.p., intraperitoneal; RMS, root mean square; s.c., subcutaneous; MJ, myoclonic jerk; PTZ, pentylenetetrazol; SD, Sprague-Dawley; SE, status epilepticus; TBI, traumatic brain injury.

4.2. TRPC5

Among the TRPC subfamily members, TRPC4 and TRPC5 proteins show the highest homology to each other, as approximately 78% of their sequences are identical (Yu, et al., 2022), underlying their similarity in gating properties. TRPC1 and TRPC4 showed the highest expression in mouse hippocampus, whereas the level of TRPC5 expression was relatively low in the same brain region (Phelan, et al., 2012). However, in mouse pilocarpine model of SE, genetic deletion of TRPC5 alone diminished the seizure severity, reduced hippocampal cell death, and largely prevented SE-associated mortality (Phelan, et al., 2013). Conversely, mice lacking both TRPC1 and TRPC4 failed to alter the seizure severity but did reduce neuronal death in hippocampal areas and lowered the overall mortality in the same SE model (Phelan, et al., 2012). These findings suggest that TRPC5 subtype likely plays more important roles in generating chemoconvulsant seizures when compared to TRPC1 and TRPC4 (Phelan, et al., 2012). Interestingly, the expression of TRPC4 and TRPC5 but not TRPC1 was elevated at both protein and mRNA levels in mouse hippocampus and cortex after traumatic brain injury (TBI) (Carver, DeWitt, Stoja, & Shapiro, 2021). Subcutaneous treatment with compound M084, an inhibitor that can block both TRPC4 and TRPC5, substantially lowered the post-TBI neuronal hyperexcitability. In addition, mice after TBI showed higher susceptibility to PTZ-induced seizures, which was also lowered by M084 (Table 2) (Carver, et al., 2021). However, whether inhibiting TRPC4 or TRPC5 alone can lead to similar outcomes remains to be determined using more selective inhibitors. As such, systemic administration of compound NU6027 in rats showed marked neuroprotection after kainate-induced SE via specifically inhibiting the gating of TRPC5 channels (Table 2) (Park, et al., 2019), although the compound was originally identified as an inhibitor of threonine-protein kinase ATR and cyclin-dependent kinase.

4.3. TRPC6

TRPC6 subtype, like TRPC3, is also extensively distributed in many regions of the brain, particularly in the hippocampus, cortex, amygdala, cerebellum, and substantia nigra (Yu, et al., 2022). However, unlike TRPC3, TRPC6 was downregulated in the hippocampus following pilocarpine-induced SE in rats (Kim, Ryu, Kim, & Kang, 2013), suggesting divergent roles of these two TRPC subfamily members in the seizing brain. Indeed, intracerebroventricular infusion of hyperforin, a natural compound that selectively activates TRPC6, protected hippocampal pyramidal neurons from SE-provoked brain cell death (Table 2), an outcome mimicking that from TRPC3 inhibition by Pyr3 (Kim, et al., 2013). In line with these findings, intracerebroventricular infusion of siRNA in rats targeting TRPC6 raised susceptibility to seizures, increased neuronal excitability, exacerbated neuronal death in dentate granule cells, but protected hippocampal cells in CA1 and CA3 regions after pilocarpine SE (Kim & Kang, 2015). However, in another study, both TRPC3 and TRPC6 were found upregulated in cortical tissues from patients with intractable TLE and in the hippocampus of mice after pilocarpine SE (Zeng, et al., 2015). Intriguingly, TRPC6 depletion by intracerebroventricular injection of an anti-TRPC6 antibody reduced dendritic arborization and spine density of hippocampal CA3 pyramidal neurons, whereas microinjection of an anti-TRPC3 antibody decreased aberrant-sprouted mossy fiber collaterals in the CA3 region of SE mice (Zeng, et al., 2015). The reasons for the contradicting expression of TRPC3 and TRPC6 in the brain after SE are unknown but might be related to different species and seizure induction protocols used in these two independent studies. Nevertheless, these two TRPC subfamily members could be differentially involved in the synaptic reorganization of mossy fiber pathway following prolonged seizures.

4.4. TRPC7

As the latest TRPC subfamily member discovered, TRPC7 is widely expressed in peripheral tissues including eye, heart, kidney, lung, intestine, and pituitary gland (Zhang & Trebak, 2014). TRPC7 is also expressed in the CNS but with functions largely unknown, and to date, there is only one study on TRPC7 channels in chemoconvulsant seizures. Congenital ablation of TRPC7 was reported to diminish the seizure induction in pilocarpine-treated mice, accompanied by reduced SE-associated mortality rate. EEG recording and analysis also revealed a decrease in gamma wave activities in TRPC7 knockout mice after pilocarpine treatment when compared to wildtype animals (Phelan, Shwe, Abramowitz, Birnbaumer, & Zheng, 2014). Further, electrophysiological study showed a reduction in epileptiform burst firing in hippocampal CA3 pyramidal neurons and the lack of high-frequency stimulation-induced long-term potentiation at CA3 and CA1 synapses in TRPC7-deficient mice. Taken together, TRPC7 might play an essential role in the initiation of acute seizures given that the epileptiform burst firing in the CA3 area is dependent on neuronal activity and considered as an essential early step of seizure generation. However, there is no pharmacological study to validate these results and determine the feasibility of TRPC7 as a therapeutic target for seizures and epilepsy.

5. TRPM channels

5.1. TRPM2

TRPM2 is widely expressed in virtually all tissues and organs, but its expression in the CNS is the highest among all TRP channels (Fonfria, et al., 2006). TRPM2 is mostly known for its important contributions to the pathogenesis of CNS conditions, such as aging, ischemic stroke, neurodegeneration, neuropathic pain, and bipolar disorder (Belrose & Jackson, 2018). However, emerging evidence suggests that TRPM2 is also involved in epileptic seizures. In an early study, TRPM2 in hippocampal neurons was found to interact with EF-hand motif-containing protein 1 (EFHC1) and contribute to the phenotypes of juvenile myoclonic epilepsy via mediating the disruptive functions of EFHC1 mutations on biological processes including neuronal apoptosis (Katano, et al., 2012). A more recent study revealed that the congenital ablation of TRPM2 in mice led to considerable antiseizure effects in PTZ, maximal electroshock, and chronic kindling models, decreased acute seizure-related neuronal death, improved cognitive functions, and alleviated brain inflammation. Importantly, the neuroprotection observed in these TRPM2 knockout mice might be related to the downregulation of PARP1/BNIP3/AIF/Endo G apoptotic pathway in cortical neurons (Zheng, et al., 2020). However, another study by the same research group showed that genetic ablation of TRPM2 in mice increased susceptibility to seizures induced by PTZ and enhanced neuronal excitability in the hippocampal CA1 region likely via inhibiting the Kv7 potassium channels (Ying, et al., 2022). The reasons for these inconsistent findings remain unclear, and future studies engaging post-SE ablation or pharmacological inhibition of TRPM2 might help to solve this contradiction.

5.2. TRPM3

TRPM3 is a heat-activated cation channel and its functions in noxious heat sensation has been well established. Like other TRP channels, the activation of TRPM3 requires phosphatidylinositol 4,5-bisphosphate. In addition to the central pore commonly found in other TRP channels, TRPM3 has an alternative pore, which shows inwardly rectifying characteristics (Zhao & Rohacs, 2021). TRPM3 is also highly expressed in the human brain (Fonfria, et al., 2006), where its roles remain largely elusive. However, recent studies reveal that mutations in TRPM3 are frequently associated with neurodevelopmental disorders, pointing to some important roles of TRPM3 in the immature brain. Particularly, a close association has been widely reported between mutations in human TRPM3 gene and the developmental and epileptic encephalopathy (DEE) (Dyment, et al., 2019; Gauthier, et al., 2021; Kang, et al., 2021), a heterogeneous group of brain disorders characterized by epilepsy and the associated intellectual disability. These DEE-associated mutations are mostly gain-of-function mutations but may have distinct effects on the TRPM3 gating via different mechanisms, including enhanced basal activity, elevated sensitivity to stimulation by high temperatures and the endogenous neurosteroid pregnenolone sulfate, as well as changed response to ligand modulation, leading to channel overactive (Van Hoeymissen, et al., 2020; Zhao, Yudin, & Rohacs, 2020). Interestingly, these mutants can be inhibited by selective TRPM3 antagonist isosakuranetin and ASD primidone, which can directly inhibit the TRPM3 ion channels (Krugel, Straub, Beckmann, & Schaefer, 2017). However, it remains to be determined whether the antiseizure effects of primidone should be attributed to its direct inhibition of TRPM3. On the other hand, it would also be interesting to know whether currently known TRPM3 inhibitors, such as isosakuranetin and liquiritigenin, have potential to be developed to treat other forms of epilepsy, besides DEE.

5.3. TRPM8

Known as the cold and menthol receptor 1, TRPM8 shows the most discrete expression patterns in human bodies among all TRP channels. Particularly, it has the highest expression in prostate, followed by liver but virtually is undetectable in other tissues and organs including the CNS (Fonfria, et al., 2006). However, TRPM8 agonist menthol has been reported to enhance tonic, but not phasic, GABAergic inhibition, suggesting a role of TRPM8 in regulating neuronal activity (Zhang, et al., 2008). Indeed, systemic treatment with M8-B, a TRPM8 antagonist that can lower the body temperature, increased the latency to febrile seizures in rat pups and led to a substantial anticonvulsant effect on PTZ-induced seizures but had no effect on electroshock-provoked convulsions (Table 3) (Zandi, Zaniani, Moghimi, & Roohbakhsh, 2019). Given that reducing body temperature can prevent febrile seizures, the antiseizure effects of M8-B is likely related to its thermoregulation, which thus may affect other temperature-sensitive TRP channels in epilepsy including TRPV4 (Shibasaki, et al., 2020). In addition, menthol was found to exert TRPM8-independent antiseizure effects in prefrontal cortex pyramidal neurons via blocking sodium channels and resembling the in vitro action of carbamazepine, a conventional ASD targeting voltage-gated sodium channels (Szulczyk & Spyrka, 2022), highlighting the limitation of using menthol to study TRPM8. As such, genetic global deletion of TRPM8 in mice exacerbated the PTZ-induced seizures and penicillin G potassium-provoked epileptiform discharges, both of which were largely suppressed by intracortical microinjection of TRPM8 agonist WS-3 in wildtype animals (Moriyama, et al., 2021). These findings together suggest that TRPM8 activation might provide a promising pharmacological strategy to manage seizures (Table 3).

Table 3.

Pharmacological studies on TRPM channels in models of seizures and epilepsy.

Pharmacological Targets Animal Models Therapeutic Agents Major Outcomes References
TRPM8 Febrile seizures induced in male Wistar rat pups M8-B (3, 6, and 9 mg/kg, i.p.) Increased the latency to the first febrile seizure Zandi et al., 2019
TRPM8 PTZ-induced seizures in male Wistar rat pups M8-B (3, 6, and 9 mg/kg, i.p.) Increased the latencies to the stage-2 and stage-4 seizures and reduced the stage-5 seizure duration Zandi et al., 2019
TRPM8 Electroshock-induced convulsions in male Wistar rat pups M8-B (3, 6, and 9 mg/kg, i.p.) Had no significant effect on convulsions Zandi et al., 2019
TRPM8 PTZ-induced seizures in male adult C57BL/6 mice WS-3 (0.1, 0.3, and 1 mM, 0.1 μL/min, i.c.) Exacerbate d PTZ-induced seizures Moriyama et al., 2021
TRPM8 PG-induced seizures in male adult C57BL/6 mice WS-3 (0.1, 0.3, and 1 mM, 0.1 μL/min, i.c.) Exacerbate d PG-induced epileptiform discharges Moriyama et al., 2021

Abbreviations: i.c., intracortical; i.p., intraperitoneal; PG, penicillin G potassium; PTZ, pentylenetetrazole.

6. Conclusions and future perspectives

6.1. What are promising?

Mounting evidence from recent studies employing both genetic and pharmacological strategies support that several members of the TRP superfamily have consistent implications in seizures of various etiologies. Among these, TRPV1, TRPV4, TRPC3, and TRPM8 likely represent the most promising TRP channels as therapeutic targets for acute seizures and potentially epilepsy. These calcium-permeable channels are often expressed at basal levels under normal conditions but can be substantially upregulated in seizing brain, particularly in the epileptic foci, demonstrating their druggability. Antiseizure effects, neuroprotection, and many other therapeutic benefits by several pharmacological compounds that target these excitatory cation channels are widely reported in various animal models, ruling out any model- or species-specific findings from these studies (Tables 13). In addition, no adverse effects of these potential therapeutic agents have been reported in these studies, indicative of their safe uses in vivo.

6.2. What are missing?

It is noted that the pharmacodynamic and pharmacokinetic profiles (e.g., potency, selectivity, in vivo half-life, and brain penetration) of some of these agents are either marginal or have not been reported (Tables 13). For instance, TRPV4-selective antagonist RN-1734 has a moderate potency (IC50: 2.3 μM) and is only about 13-fold selective against TRPV3 and TRPM8 channels (Vincent, et al., 2009). Likewise, JW-65, a recently developed TRPC3-selective inhibitor, has favorable plasma half-life (3.1 h) and brain penetration (brain-to-plasma ratio: 0.3) in mice (Nagib, et al., 2022), but it also non-negligibly blocks TRPC6 and TRPC7 channels (Zhang, et al., 2021). Future efforts in medicinal chemistry should be directed to continually improve their drug-like properties, which are essential to their development into new pharmacotherapies for seizure disorders.

In parallel to the emerging evidence from pharmacological studies, findings using various constitutive knockout mice also reveal that TRP channels have wide-ranging functions in health and disease (Wu, et al., 2010). However, global congenital deletion of TRP channels in mice could lead to some developmental adjustments and induce compensatory responses, which are quite common among some TRP subfamilies (Yu, et al., 2022). These confounding issues may underlie the controversial findings derived from studies on mice lacking TRPV1 or TRPM2. As such, in the future, adult-stage ablation or conditional, cell type-specific deletion should be used to better understand the roles of TRP channels in epilepsy, particularly, the adult-onset forms.

It is well known that seizures in human patients and experimental animals can be affected by many biological variables, such as age, weight, sex, etc., and, particularly, there is a growing recognition on sex differences in epilepsy and the associated comorbidities (Christian, Reddy, Maguire, & Forcelli, 2020). It should be pointed out that male animals (mice and rats) were exclusively used in most of these preclinical studies except one in which larval zebrafish were utilized as the model (Tables 13). This deficit should be rectified in the future studies via using both male and female animals to comply with the ARRIVE guidelines for animal research (Percie du Sert, et al., 2020) and to better understand the candidacy of TRP channels as drug targets for seizures and epilepsy by avoiding sex-biased findings.

6.3. Selectivity may not matter

It should also be noted that some of the TRP superfamily members can form heteromeric complex channels with others, and this is especially true in certain brain areas. For example, TRPC3 isomers alone can polymerize to form operational homomeric channels, but they can also multimerize with other subtypes of TRPC, particularly TRPC6 and TRPC7 due to their similarities in amino acid sequence, three-dimensional structure, signal transduction, and expression pattern (Yu, et al., 2022). The heteromerization among different TRP channel members can be cell type-specific, function-dependent, and disease-related (Zhang & Trebak, 2014). It’s unclear whether TRP homomultimers or heteromultimers are chiefly responsible for the pathogenic processes in the epileptic brain. If culprit is the latter, targeting a specific TRP member by highly selective compounds may not provide satisfactory therapeutic benefits. This possibility and the potential compensation among the TRP superfamily members, which is actually quite common (Wang, et al., 2020), would explain, at least partially, why the genetic deletion of a single TRP channel (e.g., TRPC3) was not able to fully terminate seizures in most of these studies. As such, less selective compounds that can act on multiple TRP superfamily members might have advantages and should not be abandoned for evaluations in animal seizure models.

6.4. Multiple roads to treatment

Most pharmacological studies to date have been specifically focusing on the effects of these TRP-targeting compounds on acute seizures. Future research should also be directed to evaluate their effects on the development of spontaneous recurrent seizures after precipitating brain insults, such as TBI, stroke, brain infection, de novo SE, brain tumor, etc. Outcomes will help to determine whether targeting TRP channels can lead to the prevention of acquired epilepsy (i.e., epileptogenesis), modification of the disease progression, or even a cure, the ultimate goal in epilepsy management that no current FDA-approved drugs have been demonstrated to be able to achieve (Chen, et al., 2023; Jiang, Santhakumar, & Zhu, 2022; Loscher, 2020). Another crucial but unsolved issue that substantially impedes the current antiseizure medication is pharmacoresistance, which is commonly observed in more than 30% patients with epilepsy (Janmohamed, et al., 2020). Therefore, it is also important to test these TRP-targeting compounds in preclinical models of refractory epilepsy and to determine whether they can enhance the sensitivity to current ASDs when seizures become pharmacoresistant (Löscher, 2017).

Acknowledgments

This work was supported by the National Institutes of Health (NIH)/National Institute of Neurological Disorders and Stroke (NINDS) grants R01NS100947 (J.J.), R21NS109687 (J.J.), and R61NS124923 (J.J.). The content is solely the responsibility of the authors and does not necessarily represent the official views of the NIH.

Abbreviations

AIF

apoptosis-inducing factor

4-AP

4-aminopyridine

ARRIVE

Animal Research: Reporting of In Vivo Experiments

ASD

antiseizure drug

BNIP3

BCL2/adenovirus E1B 19 kDa protein-interacting protein 3

CNS

central nervous system

DEE

developmental and epileptic encephalopathy

EEG

electroencephalogram

EFHC1

EF-hand motif-containing protein 1

Endo G

endonuclease G

GABA

γ-aminobutyric acid

GFAP

glial fibrillary acidic protein

HLA-DR

human leukocyte antigen - DR isotype

Iba1

ionized calcium-binding adapter molecule 1

IL-1β

interleukin 1β

IL-6

interleukin 6

KCHIP

potassium channel interacting protein

MDA

maximal dentate activation

NeuN

neuronal nuclear antigen

NMDA

N-methyl-D-aspartate

OLDA

N-oleoyldopamine

PARP1

poly (ADP-ribose) polymerase 1

PBCH

N-phenyl-N’-(4-benzyloxyphenoxycarbonyl)-4-chlorophenylsulfonyl hydrazide

PTZ

pentylenetetrazole

SE

status epilepticus

TBI

traumatic brain injury

TLE

temporal lobe epilepsy

TNF-α

tumor necrosis factor α

TRP

transient receptor potential channels

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

Conflict of Interest Statement

The authors declare that there are no conflicts of interest.

Data Availability

No data was used for the research described in the article.

References

  1. Belrose JC, & Jackson MF (2018). TRPM2: a candidate therapeutic target for treating neurological diseases. Acta Pharmacol Sin 39, 722–732. [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Brini M, Cali T, Ottolini D, & Carafoli E (2014). Neuronal calcium signaling: function and dysfunction. Cell Mol Life Sci 71, 2787–2814. [DOI] [PMC free article] [PubMed] [Google Scholar]
  3. Carletti F, Gambino G, Rizzo V, Ferraro G, & Sardo P (2016). Involvement of TRPV1 channels in the activity of the cannabinoid WIN 55,212–2 in an acute rat model of temporal lobe epilepsy. Epilepsy Res 122, 56–65. [DOI] [PubMed] [Google Scholar]
  4. Carver CM, DeWitt HR, Stoja AP, & Shapiro MS (2021). Blockade of TRPC Channels Limits Cholinergic-Driven Hyperexcitability and Seizure Susceptibility After Traumatic Brain Injury. Front Neurosci 15, 681144. [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Carvill GL, Dulla CG, Lowenstein DH, & Brooks-Kayal AR (2020). The path from scientific discovery to cures for epilepsy. Neuropharmacology 167, 107702. [DOI] [PubMed] [Google Scholar]
  6. Chen Y, Nagib MM, Yasmen N, Sluter MN, Littlejohn TL, Yu Y, & Jiang J (2023). Neuroinflammatory mediators in acquired epilepsy: an update. Inflamm Res. [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Chen Z, Brodie MJ, Liew D, & Kwan P (2018). Treatment Outcomes in Patients With Newly Diagnosed Epilepsy Treated With Established and New Antiepileptic Drugs: A 30-Year Longitudinal Cohort Study. JAMA Neurol 75, 279–286. [DOI] [PMC free article] [PubMed] [Google Scholar]
  8. Christian CA, Reddy DS, Maguire J, & Forcelli PA (2020). Sex Differences in the Epilepsies and Associated Comorbidities: Implications for Use and Development of Pharmacotherapies. Pharmacol Rev 72, 767–800. [DOI] [PMC free article] [PubMed] [Google Scholar]
  9. Devinsky O, Vezzani A, O’Brien TJ, Jette N, Scheffer IE, de Curtis M, & Perucca P (2018). Epilepsy. Nat Rev Dis Primers 4, 18024. [DOI] [PubMed] [Google Scholar]
  10. Dey A, Kang X, Qiu J, Du Y, & Jiang J (2016). Anti-Inflammatory Small Molecules To Treat Seizures and Epilepsy: From Bench to Bedside. Trends Pharmacol Sci 37, 463–484. [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Dyment DA, Terhal PA, Rustad CF, Tveten K, Griffith C, Jayakar P, Shinawi M, Ellingwood S, Smith R, van Gassen K, McWalter K, Innes AM, & Lines MA (2019). De novo substitutions of TRPM3 cause intellectual disability and epilepsy. Eur J Hum Genet 27, 1611–1618. [DOI] [PMC free article] [PubMed] [Google Scholar]
  12. Fisher RS, Cross JH, D’Souza C, French JA, Haut SR, Higurashi N, Hirsch E, Jansen FE, Lagae L, Moshe SL, Peltola J, Roulet Perez E, Scheffer IE, Schulze-Bonhage A, Somerville E, Sperling M, Yacubian EM, & Zuberi SM (2017a). Instruction manual for the ILAE 2017 operational classification of seizure types. Epilepsia 58, 531–542. [DOI] [PubMed] [Google Scholar]
  13. Fisher RS, Cross JH, French JA, Higurashi N, Hirsch E, Jansen FE, Lagae L, Moshe SL, Peltola J, Roulet Perez E, Scheffer IE, & Zuberi SM (2017b). Operational classification of seizure types by the International League Against Epilepsy: Position Paper of the ILAE Commission for Classification and Terminology. Epilepsia 58, 522–530. [DOI] [PubMed] [Google Scholar]
  14. Fonfria E, Murdock PR, Cusdin FS, Benham CD, Kelsell RE, & McNulty S (2006). Tissue distribution profiles of the human TRPM cation channel family. J Recept Signal Transduct Res 26, 159–178. [DOI] [PubMed] [Google Scholar]
  15. Galanopoulou AS, Loscher W, Lubbers L, O’Brien TJ, Staley K, Vezzani A, D’Ambrosio R, White HS, Sontheimer H, Wolf JA, Twyman R, Whittemore V, Wilcox KS, & Klein B (2021). Antiepileptogenesis and disease modification: Progress, challenges, and the path forward-Report of the Preclinical Working Group of the 2018 NINDS-sponsored antiepileptogenesis and disease modification workshop. Epilepsia Open 6, 276–296. [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Garcia-Rodriguez C, Bravo-Tobar ID, Duarte Y, Barrio LC, & Saez JC (2022). Contribution of non-selective membrane channels and receptors in epilepsy. Pharmacol Ther 231, 107980. [DOI] [PubMed] [Google Scholar]
  17. Gauthier LW, Chatron N, Cabet S, Labalme A, Carneiro M, Poirot I, Delvert C, Gleizal A, Lesca G, & Putoux A (2021). Description of a novel patient with the TRPM3 recurrent p.Val837Met variant. Eur J Med Genet 64, 104320. [DOI] [PubMed] [Google Scholar]
  18. Gladkikh IN, Sintsova OV, Leychenko EV, & Kozlov SA (2021). TRPV1 Ion Channel: Structural Features, Activity Modulators, and Therapeutic Potential. Biochemistry (Mosc) 86, S50–S70. [DOI] [PubMed] [Google Scholar]
  19. Harteneck C, & Leuner K (2014). TRP Channels in Neuronal and Glial Signal Transduction. In Thomas H (Ed.), Neurochemistry (pp. Ch. 2). Rijeka: IntechOpen. [Google Scholar]
  20. Hartmann J, Dragicevic E, Adelsberger H, Henning HA, Sumser M, Abramowitz J, Blum R, Dietrich A, Freichel M, Flockerzi V, Birnbaumer L, & Konnerth A (2008). TRPC3 channels are required for synaptic transmission and motor coordination. Neuron 59, 392–398. [DOI] [PMC free article] [PubMed] [Google Scholar]
  21. Hauser WA (2018). Questioning the Effectiveness of Newer Antiseizure Medications. JAMA Neurol 75, 273–274. [DOI] [PubMed] [Google Scholar]
  22. Hendrickx DAE, van Eden CG, Schuurman KG, Hamann J, & Huitinga I (2017). Staining of HLA-DR, Iba1 and CD68 in human microglia reveals partially overlapping expression depending on cellular morphology and pathology. J Neuroimmunol 309, 12–22. [DOI] [PubMed] [Google Scholar]
  23. Heuser K, Nome CG, Pettersen KH, Abjorsbraten KS, Jensen V, Tang W, Sprengel R, Tauboll E, Nagelhus EA, & Enger R (2018). Ca2+ Signals in Astrocytes Facilitate Spread of Epileptiform Activity. Cereb Cortex 28, 4036–4048. [DOI] [PMC free article] [PubMed] [Google Scholar]
  24. Hunt RF, Hortopan GA, Gillespie A, & Baraban SC (2012). A novel zebrafish model of hyperthermia-induced seizures reveals a role for TRPV4 channels and NMDA-type glutamate receptors. Exp Neurol 237, 199–206. [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. Janmohamed M, Brodie MJ, & Kwan P (2020). Pharmacoresistance - Epidemiology, mechanisms, and impact on epilepsy treatment. Neuropharmacology 168, 107790. [DOI] [PubMed] [Google Scholar]
  26. Jia YF, Li YC, Tang YP, Cao J, Wang LP, Yang YX, Xu L, & Mao RR (2015). Interference of TRPV1 function altered the susceptibility of PTZ-induced seizures. Front Cell Neurosci 9, 20. [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. Jiang J, Santhakumar V, & Zhu X (2022). Editorial: Neuroinflammation in acquired epilepsy. Front Cell Dev Biol 10, 1074537. [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Kang Q, Yang L, Liao H, Yang S, Kuang X, Ning Z, Liao C, & Chen B (2021). A Chinese patient with developmental and epileptic encephalopathies (DEE) carrying a TRPM3 gene mutation: a paediatric case report. BMC Pediatr 21, 256. [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. Katano M, Numata T, Aguan K, Hara Y, Kiyonaka S, Yamamoto S, Miki T, Sawamura S, Suzuki T, Yamakawa K, & Mori Y (2012). The juvenile myoclonic epilepsy-related protein EFHC1 interacts with the redox-sensitive TRPM2 channel linked to cell death. Cell Calcium 51, 179–185. [DOI] [PubMed] [Google Scholar]
  30. Kawamoto EM, Vivar C, & Camandola S (2012). Physiology and pathology of calcium signaling in the brain. Front Pharmacol 3, 61. [DOI] [PMC free article] [PubMed] [Google Scholar]
  31. Kim DS, Ryu HJ, Kim JE, & Kang TC (2013). The reverse roles of transient receptor potential canonical channel-3 and −6 in neuronal death following pilocarpine-induced status epilepticus. Cell Mol Neurobiol 33, 99–109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  32. Kim YJ, & Kang TC (2015). The role of TRPC6 in seizure susceptibility and seizure-related neuronal damage in the rat dentate gyrus. Neuroscience 307, 215–230. [DOI] [PubMed] [Google Scholar]
  33. Kiyonaka S, Kato K, Nishida M, Mio K, Numaga T, Sawaguchi Y, Yoshida T, Wakamori M, Mori E, Numata T, Ishii M, Takemoto H, Ojida A, Watanabe K, Uemura A, Kurose H, Morii T, Kobayashi T, Sato Y, Sato C, Hamachi I, & Mori Y (2009). Selective and direct inhibition of TRPC3 channels underlies biological activities of a pyrazole compound. Proc Natl Acad Sci U S A 106, 5400–5405. [DOI] [PMC free article] [PubMed] [Google Scholar]
  34. Koivisto AP, Belvisi MG, Gaudet R, & Szallasi A (2022). Advances in TRP channel drug discovery: from target validation to clinical studies. Nat Rev Drug Discov 21, 41–59. [DOI] [PMC free article] [PubMed] [Google Scholar]
  35. Kong W, Wang X, Yang X, Huang W, Han S, Yin J, Liu W, He X, & Peng B (2019). Activation of TRPV1 Contributes to Recurrent Febrile Seizures via Inhibiting the Microglial M2 Phenotype in the Immature Brain. Front Cell Neurosci 13, 442. [DOI] [PMC free article] [PubMed] [Google Scholar]
  36. Kong WL, Min JW, Liu YL, Li JX, He XH, & Peng BW (2014). Role of TRPV1 in susceptibility to PTZ-induced seizure following repeated hyperthermia challenges in neonatal mice. Epilepsy Behav 31, 276–280. [DOI] [PubMed] [Google Scholar]
  37. Krugel U, Straub I, Beckmann H, & Schaefer M (2017). Primidone inhibits TRPM3 and attenuates thermal nociception in vivo. Pain 158, 856–867. [DOI] [PMC free article] [PubMed] [Google Scholar]
  38. Li Y, Calfa G, Inoue T, Amaral MD, & Pozzo-Miller L (2010). Activity-dependent release of endogenous BDNF from mossy fibers evokes a TRPC3 current and Ca2+ elevations in CA3 pyramidal neurons. J Neurophysiol 103, 2846–2856. [DOI] [PMC free article] [PubMed] [Google Scholar]
  39. Loscher W (2020). The holy grail of epilepsy prevention: Preclinical approaches to antiepileptogenic treatments. Neuropharmacology 167, 107605. [DOI] [PubMed] [Google Scholar]
  40. Löscher W (2017). Animal Models of Drug-Refractory Epilepsy. In Pitkänen A, Buckmaster PS, Galanopoulou AS & Moshé SL (Eds.), Models of Seizures and Epilepsy (pp. 743–760): Academic Press. [Google Scholar]
  41. Loscher W, & Klein P (2020). The feast and famine: Epilepsy treatment and treatment gaps in early 21st century. Neuropharmacology 170, 108055. [DOI] [PubMed] [Google Scholar]
  42. Men C, Wang Z, Zhou L, Qi M, An D, Xu W, Zhan Y, & Chen L (2019). Transient receptor potential vanilloid 4 is involved in the upregulation of connexin expression following pilocarpine-induced status epilepticus in mice. Brain Res Bull 152, 128–133. [DOI] [PubMed] [Google Scholar]
  43. Meza RC, Ancaten-Gonzalez C, Chiu CQ, & Chavez AE (2022). Transient Receptor Potential Vanilloid 1 Function at Central Synapses in Health and Disease. Front Cell Neurosci 16, 864828. [DOI] [PMC free article] [PubMed] [Google Scholar]
  44. Mitsumura K, Hosoi N, Furuya N, & Hirai H (2011). Disruption of metabotropic glutamate receptor signalling is a major defect at cerebellar parallel fibre-Purkinje cell synapses in staggerer mutant mice. J Physiol 589, 3191–3209. [DOI] [PMC free article] [PubMed] [Google Scholar]
  45. Moriyama H, Nomura S, Imoto H, Inoue T, Fujiyama Y, Haji K, Maruta Y, Ishihara H, & Suzuki M (2021). Suppressive Effects of Transient Receptor Potential Melastatin 8 Agonist on Epileptiform Discharges and Epileptic Seizures. Front Pharmacol 12, 766782. [DOI] [PMC free article] [PubMed] [Google Scholar]
  46. Nagib MM, Zhang S, Yasmen N, Li L, Hou R, Yu Y, Boda VK, Wu Z, Li W, & Jiang J (2022). Inhibition of TRPC3 channels by a novel pyrazole compound confers antiseizure effects. Epilepsia 63, 1003–1015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  47. Nilius B, & Szallasi A (2014). Transient receptor potential channels as drug targets: from the science of basic research to the art of medicine. Pharmacol Rev 66, 676–814. [DOI] [PubMed] [Google Scholar]
  48. Park SE, Song JH, Hong C, Kim DE, Sul JW, Kim TY, Seo BR, So I, Kim SY, Bae DJ, Park MH, Lim HM, Baek IJ, Riccio A, Lee JY, Shim WH, Park B, Koh JY, & Hwang JJ (2019). Contribution of Zinc-Dependent Delayed Calcium Influx via TRPC5 in Oxidative Neuronal Death and its Prevention by Novel TRPC Antagonist. Mol Neurobiol 56, 2822–2835. [DOI] [PMC free article] [PubMed] [Google Scholar]
  49. Penn Y, Segal M, & Moses E (2016). Network synchronization in hippocampal neurons. Proc Natl Acad Sci U S A 113, 3341–3346. [DOI] [PMC free article] [PubMed] [Google Scholar]
  50. Percie du Sert N, Ahluwalia A, Alam S, Avey MT, Baker M, Browne WJ, Clark A, Cuthill IC, Dirnagl U, Emerson M, Garner P, Holgate ST, Howells DW, Hurst V, Karp NA, Lazic SE, Lidster K, MacCallum CJ, Macleod M, Pearl EJ, Petersen OH, Rawle F, Reynolds P, Rooney K, Sena ES, Silberberg SD, Steckler T, & Wurbel H (2020). Reporting animal research: Explanation and elaboration for the ARRIVE guidelines 2.0. PLoS Biol 18, e3000411. [DOI] [PMC free article] [PubMed] [Google Scholar]
  51. Phelan KD, Mock MM, Kretz O, Shwe UT, Kozhemyakin M, Greenfield LJ, Dietrich A, Birnbaumer L, Freichel M, Flockerzi V, & Zheng F (2012). Heteromeric canonical transient receptor potential 1 and 4 channels play a critical role in epileptiform burst firing and seizure-induced neurodegeneration. Mol Pharmacol 81, 384–392. [DOI] [PMC free article] [PubMed] [Google Scholar]
  52. Phelan KD, Shwe UT, Abramowitz J, Birnbaumer L, & Zheng F (2014). Critical role of canonical transient receptor potential channel 7 in initiation of seizures. Proc Natl Acad Sci U S A 111, 11533–11538. [DOI] [PMC free article] [PubMed] [Google Scholar]
  53. Phelan KD, Shwe UT, Abramowitz J, Wu H, Rhee SW, Howell MD, Gottschall PE, Freichel M, Flockerzi V, Birnbaumer L, & Zheng F (2013). Canonical transient receptor channel 5 (TRPC5) and TRPC1/4 contribute to seizure and excitotoxicity by distinct cellular mechanisms. Mol Pharmacol 83, 429–438. [DOI] [PMC free article] [PubMed] [Google Scholar]
  54. Phelan KD, Shwe UT, Cozart MA, Wu H, Mock MM, Abramowitz J, Birnbaumer L, & Zheng F (2017). TRPC3 channels play a critical role in the theta component of pilocarpine-induced status epilepticus in mice. Epilepsia 58, 247–254. [DOI] [PMC free article] [PubMed] [Google Scholar]
  55. Prakriya M, & Lewis RS (2015). Store-Operated Calcium Channels. Physiol Rev 95, 1383–1436. [DOI] [PMC free article] [PubMed] [Google Scholar]
  56. Rho JM, & White HS (2018). Brief history of anti-seizure drug development. Epilepsia Open 3, 114–119. [DOI] [PMC free article] [PubMed] [Google Scholar]
  57. Roedding AS, Gao AF, Wu AM, Li PP, Kish SJ, & Warsh JJ (2009). TRPC3 protein is expressed across the lifespan in human prefrontal cortex and cerebellum. Brain Res 1260, 1–6. [DOI] [PubMed] [Google Scholar]
  58. Samanta A, Hughes TET, & Moiseenkova-Bell VY (2018). Transient Receptor Potential (TRP) Channels. Subcell Biochem 87, 141–165. [DOI] [PMC free article] [PubMed] [Google Scholar]
  59. Sasaki T, Ishikawa T, Abe R, Nakayama R, Asada A, Matsuki N, & Ikegaya Y (2014). Astrocyte calcium signalling orchestrates neuronal synchronization in organotypic hippocampal slices. J Physiol 592, 2771–2783. [DOI] [PMC free article] [PubMed] [Google Scholar]
  60. Shibasaki K, Ikenaka K, Tamalu F, Tominaga M, & Ishizaki Y (2014). A novel subtype of astrocytes expressing TRPV4 (transient receptor potential vanilloid 4) regulates neuronal excitability via release of gliotransmitters. J Biol Chem 289, 14470–14480. [DOI] [PMC free article] [PubMed] [Google Scholar]
  61. Shibasaki K, Yamada K, Miwa H, Yanagawa Y, Suzuki M, Tominaga M, & Ishizaki Y (2020). Temperature elevation in epileptogenic foci exacerbates epileptic discharge through TRPV4 activation. Lab Invest 100, 274–284. [DOI] [PubMed] [Google Scholar]
  62. Shirazi M, Izadi M, Amin M, Rezvani ME, Roohbakhsh A, & Shamsizadeh A (2014). Involvement of central TRPV1 receptors in pentylenetetrazole and amygdala-induced kindling in male rats. Neurol Sci 35, 1235–1241. [DOI] [PubMed] [Google Scholar]
  63. Sills GJ, & Rogawski MA (2020). Mechanisms of action of currently used antiseizure drugs. Neuropharmacology 168, 107966. [DOI] [PubMed] [Google Scholar]
  64. Suemaru K, Yoshikawa M, Aso H, & Watanabe M (2018). TRPV1 mediates the anticonvulsant effects of acetaminophen in mice. Epilepsy Res 145, 153–159. [DOI] [PubMed] [Google Scholar]
  65. Sun D, Ma H, Ma J, Wang J, Deng X, Hu C, & Deng X (2018). Canonical Transient Receptor Potential Channel 3 Contributes to Febrile Seizure Inducing Neuronal Cell Death and Neuroinflammation. Cell Mol Neurobiol 38, 1215–1226. [DOI] [PMC free article] [PubMed] [Google Scholar]
  66. Sun FJ, Guo W, Zheng DH, Zhang CQ, Li S, Liu SY, Yin Q, Yang H, & Shu HF (2013). Increased expression of TRPV1 in the cortex and hippocampus from patients with mesial temporal lobe epilepsy. J Mol Neurosci 49, 182–193. [DOI] [PubMed] [Google Scholar]
  67. Szulczyk B, & Spyrka A (2022). Menthol exerts TRPM8-independent antiepileptic effects in prefrontal cortex pyramidal neurons. Brain Res 1783, 147847. [DOI] [PubMed] [Google Scholar]
  68. Tan X, Zeng Y, Tu Z, Li P, Chen H, Cheng L, Tu S, & Jiang L (2020). TRPV1 Contributes to the Neuroprotective Effect of Dexmedetomidine in Pilocarpine-Induced Status Epilepticus Juvenile Rats. Biomed Res Int 2020, 7623635. [DOI] [PMC free article] [PubMed] [Google Scholar]
  69. Van Hoeymissen E, Held K, Nogueira Freitas AC, Janssens A, Voets T, & Vriens J (2020). Gain of channel function and modified gating properties in TRPM3 mutants causing intellectual disability and epilepsy. Elife 9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  70. Varvel NH, Jiang J, & Dingledine R (2015). Candidate drug targets for prevention or modification of epilepsy. Annu Rev Pharmacol Toxicol 55, 229–247. [DOI] [PMC free article] [PubMed] [Google Scholar]
  71. Vincent F, Acevedo A, Nguyen MT, Dourado M, DeFalco J, Gustafson A, Spiro P, Emerling DE, Kelly MG, & Duncton MA (2009). Identification and characterization of novel TRPV4 modulators. Biochem Biophys Res Commun 389, 490–494. [DOI] [PubMed] [Google Scholar]
  72. Wang H, Cheng X, Tian J, Xiao Y, Tian T, Xu F, Hong X, & Zhu MX (2020). TRPC channels: Structure, function, regulation and recent advances in small molecular probes. Pharmacol Ther 209, 107497. [DOI] [PMC free article] [PubMed] [Google Scholar]
  73. Wang Z, Zhou L, An D, Xu W, Wu C, Sha S, Li Y, Zhu Y, Chen A, Du Y, Chen L, & Chen L (2019). TRPV4-induced inflammatory response is involved in neuronal death in pilocarpine model of temporal lobe epilepsy in mice. Cell Death Dis 10, 386. [DOI] [PMC free article] [PubMed] [Google Scholar]
  74. Wu LJ, Sweet TB, & Clapham DE (2010). International Union of Basic and Clinical Pharmacology. LXXVI. Current progress in the mammalian TRP ion channel family. Pharmacol Rev 62, 381–404. [DOI] [PMC free article] [PubMed] [Google Scholar]
  75. Xu W, Wang Y, Qi X, Li K, Zhou L, Sha S, Wang X, Wu C, Du Y, & Chen L (2022). Involvement of TRPV4 in changes in rapidly inactivating potassium channels in the early stage of pilocarpine-induced status epilepticus in mice. J Cell Physiol 237, 856–867. [DOI] [PubMed] [Google Scholar]
  76. Yasmen N, Sluter MN, Yu Y, & Jiang J (2023). Ganaxolone for management of seizures associated with CDKL5 deficiency disorder. Trends Pharmacol Sci 44, 128–129. [DOI] [PMC free article] [PubMed] [Google Scholar]
  77. Ying Y, Gong L, Tao X, Ding J, Chen N, Yao Y, Liu J, Chen C, Zhu T, & Jiang P (2022). Genetic Knockout of TRPM2 Increases Neuronal Excitability of Hippocampal Neurons by Inhibiting Kv7 Channel in Epilepsy. Mol Neurobiol 59, 6918–6933. [DOI] [PubMed] [Google Scholar]
  78. Yu Y, Li W, & Jiang J (2022). TRPC channels as emerging targets for seizure disorders. Trends Pharmacol Sci 43, 787–798. [DOI] [PMC free article] [PubMed] [Google Scholar]
  79. Yu Y, Nguyen DT, & Jiang J (2019). G protein-coupled receptors in acquired epilepsy: Druggability and translatability. Prog Neurobiol 183, 101682. [DOI] [PMC free article] [PubMed] [Google Scholar]
  80. Zandi N, Zaniani NR, Moghimi A, & Roohbakhsh A (2019). Protective effects of M8-B, a TRPM8 antagonist, on febrile- and pentylenetetrazol-induced seizures. Acta Neurobiol Exp (Wars) 79, 86–91. [PubMed] [Google Scholar]
  81. Zeng C, Zhou P, Jiang T, Yuan C, Ma Y, Feng L, Liu R, Tang W, Long X, Xiao B, & Tian F (2015). Upregulation and Diverse Roles of TRPC3 and TRPC6 in Synaptic Reorganization of the Mossy Fiber Pathway in Temporal Lobe Epilepsy. Mol Neurobiol 52, 562–572. [DOI] [PubMed] [Google Scholar]
  82. Zeng ML, Cheng JJ, Kong S, Yang XL, Jia XL, Cheng XL, Chen L, He FG, Liu YM, Fan YT, Gongga L, Chen TX, Liu WH, He XH, & Peng BW (2022). Inhibition of Transient Receptor Potential Vanilloid 4 (TRPV4) Mitigates Seizures. Neurotherapeutics 19, 660–681. [DOI] [PMC free article] [PubMed] [Google Scholar]
  83. Zeng ML, Kong S, Chen TX, & Peng BW (2023). Transient Receptor Potential Vanilloid 4: a Double-Edged Sword in the Central Nervous System. Mol Neurobiol 60, 1232–1249. [DOI] [PubMed] [Google Scholar]
  84. Zhang S, Romero LO, Deng S, Wang J, Li Y, Yang L, Hamilton DJ, Miller DD, Liao FF, Cordero-Morales JF, Wu Z, & Li W (2021). Discovery of a Highly Selective and Potent TRPC3 Inhibitor with High Metabolic Stability and Low Toxicity. ACS Med Chem Lett 12, 572–578. [DOI] [PMC free article] [PubMed] [Google Scholar]
  85. Zhang X, & Trebak M (2014). Transient receptor potential canonical 7: a diacylglycerol-activated non-selective cation channel. Handb Exp Pharmacol 222, 189–204. [DOI] [PMC free article] [PubMed] [Google Scholar]
  86. Zhang XB, Jiang P, Gong N, Hu XL, Fei D, Xiong ZQ, Xu L, & Xu TL (2008). A-type GABA receptor as a central target of TRPM8 agonist menthol. PLoS One 3, e3386. [DOI] [PMC free article] [PubMed] [Google Scholar]
  87. Zhao S, & Rohacs T (2021). The newest TRP channelopathy: Gain of function TRPM3 mutations cause epilepsy and intellectual disability. Channels (Austin) 15, 386–397. [DOI] [PMC free article] [PubMed] [Google Scholar]
  88. Zhao S, Yudin Y, & Rohacs T (2020). Disease-associated mutations in the human TRPM3 render the channel overactive via two distinct mechanisms. Elife 9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  89. Zheng Q, Zhu T, Hu H, Zhao Y, Ying Y, Luo X, Ling Y, Chen Z, Ji H, & Jiang P (2020). TRPM2 ion channel is involved in the aggravation of cognitive impairment and down regulation of epilepsy threshold in pentylenetetrazole-induced kindling mice. Brain Res Bull 155, 48–60. [DOI] [PubMed] [Google Scholar]
  90. Zhou FM, & Lee CR (2011). Intrinsic and integrative properties of substantia nigra pars reticulata neurons. Neuroscience 198, 69–94. [DOI] [PMC free article] [PubMed] [Google Scholar]

Associated Data

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

Data Availability Statement

No data was used for the research described in the article.

RESOURCES