Next Article in Journal
Operation of Hybrid Membranes for the Removal of Pharmaceuticals and Pollutants from Water and Wastewater
Next Article in Special Issue
Interactions between the Nicotinic and Endocannabinoid Receptors at the Plasma Membrane
Previous Article in Journal
Progress in the Degradability of Biodegradable Film Materials for Packaging
Previous Article in Special Issue
Understanding the Dynamics of the Transient and Permanent Opening Events of the Mitochondrial Permeability Transition Pore with a Novel Stochastic Model
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Behavior of KCNQ Channels in Neural Plasticity and Motor Disorders

Department of Biomedical Sciences, School of Medicine, University of Missouri-Kansas City, Kansas City, MO 64108, USA
*
Author to whom correspondence should be addressed.
Membranes 2022, 12(5), 499; https://doi.org/10.3390/membranes12050499
Submission received: 29 March 2022 / Revised: 26 April 2022 / Accepted: 3 May 2022 / Published: 6 May 2022
(This article belongs to the Special Issue Membrane Channel of Cells)

Abstract

:
The broad distribution of voltage-gated potassium channels (VGKCs) in the human body makes them a critical component for the study of physiological and pathological function. Within the KCNQ family of VGKCs, these aqueous conduits serve an array of critical roles in homeostasis, especially in neural tissue. Moreover, the greater emphasis on genomic identification in the past century has led to a growth in literature on the role of the ion channels in pathological disease as well. Despite this, there is a need to consolidate the updated findings regarding both the pharmacotherapeutic and pathological roles of KCNQ channels, especially regarding neural plasticity and motor disorders which have the largest body of literature on this channel. Specifically, KCNQ channels serve a remarkable role in modulating the synaptic efficiency required to create appropriate plasticity in the brain. This role can serve as a foundation for clinical approaches to chronic pain. Additionally, KCNQ channels in motor disorders have been utilized as a direction for contemporary pharmacotherapeutic developments due to the muscarinic properties of this channel. The aim of this study is to provide a contemporary review of the behavior of these channels in neural plasticity and motor disorders. Upon review, the behavior of these channels is largely dependent on the physiological role that KCNQ modulatory factors (i.e., pharmacotherapeutic options) serve in pathological diseases.

1. Introduction

Ion channels serve as an aqueous conduit for several nuanced cellular processes to maintain the homeostatic direction of the body. Moreover, there are over 400 genes that encode for at least one ion channel subunit [1,2]. The various mechanisms for alternative splicing make for an enormous variety of subunit combinations designed for appropriate physiological functions. Among these, the largest and most diverse group of ion channels are potassium (K+) channels [2,3]. These channels are composed of tetrameric integral membrane regions, which form an aqueous pore for K+ to permeate across the membrane. This ion serves a critical role in maintaining electrical gradients during the repolarization of action potentials and maintaining the negative resting membrane potential [3,4].
Voltage-gated potassium channels (VGKCs, also Kv) form a broad distribution of channels in the nervous system as well as other tissues. Structurally, Kv channels are also a tetramer integral membrane pore-forming alpha subunit but also contain six transmembrane segmental helices, classified as S1–S6. In addition, the S1–S4 transmembrane segmental helices compose the actual voltage sensation region, and the latter two (S5–S6) units are the actual gate of the channel, as depicted in Figure 1. The voltage sensation region (S1–S4) is supple in its ability to adapt to shifting membrane potentials by creating a conformational shift. This shift spreads through the pore-forming subunit via interactions with the S4 transmembrane segments. In addition, this segment is also protected during depolarization of the action potential (AP). This protection is due to the presence of the acidic residues on S1 and S2 transmembrane segments, which limits deterrence [3,4,5].
Within the family of Kv channels, there are subfamilies that can be grouped according to the N- and C-terminal domains and encoded genes [5,6]. The importance behind the subfamily grouping lies in the Kv proteins, which can be functionally divergent with different membrane sensitivity potentials, gating interactions, and dynamic responses [4]. These subfamilies of Kv channels are all encoded by 40 genes, and current literature establishes exactly 12 subfamilies of Kv channels as a product of this gene encoding (e.g., Kv1–12) [6].
Historically, some of the earliest studies on voltage-gated ion channels (VGICs) were on the contemporary Kv7 subfamily [5,6]. Moreover, the understanding of the Kv7 subfamily was not immediate upon discovery. Rather, the literature initially focused on a concept known as the M channel. This channel was initially termed due to its activity as a low-threshold non-inactivating K+ channel [7]. They were named “M channels” as such because of pilot literature that showcased their inhibition via muscarinic acetylcholine receptors (mAChR) stimulation [5]. Today, the subunits of the subfamily Q Kv7 K+ (KCNQ) channel family are now known to be part of M channels and are a key target as the basis for pharmacological treatment modalities for a broad spectrum of neurological disorders. This is because Kv7 have been shown to be stimulated by membrane potentials that are more negative than the AP threshold due to their activity as a low-threshold non-inactivating K+ channel [5,6,7].
Structurally, the KCNQ channels are similar to their Kv channel relatives (Figure 1). However, the emphasis on these channels is in their ability to utilize their glycine residues to contribute to a major part of their K+ ion preference [8,9]. Specifically, the channels have glycine residues which utilize their carbonyl oxygen branches to form a shell that is specific for the size of K+ ions compared to Ca2+ and Na+ ions [9,10].
The KCNQ channels are responsible for the M currents during physiological processes, which is important in the regulation of various neuronal excitability [10]. The basis of which is formed by several different KCNQ isoforms forming heterotrimeric channels. The M-current is a non-inactivating sub-threshold current [9,10]. The increases in neuronal excitability have resulted from physiological modulation, pharmacological inhibition, and genetic mutations that affect the M-current [9,10,11]. The Kv7 channels can transiently induce the suppression of the M-current such that they limit the firing frequency of neurons [10,11]. Furthermore, it is the Kv7.2 and Kv7.3 channels which are specifically involved in the regulation of M-current, and some other channels can also play minor contributory roles [10,11,12,13].
With regards to the actual opening and closing of the KCNQ channel, there are several mechanisms. For example, KCNQ channels can open via binding of the phosphatidylinositol 4,5-bisphosphate (PIP2) ligand. The direct binding of gamma-aminobutyric acid (GABA) to the KCNQ channel can directly increase the likelihood that a KCNQ channel will open and allow K+ permeation. This mechanism seems to be GABA-specific as such a conformation has not been identified in KCNQ channels activated by other means. Secondly, inositol 1,4,5-trisphosphate (IP3)-mediated intracellular calcium signals promote PIP2 synthesis and, via calmodulin, will suppress the M-current [14,15]. In regard to neuronal KCNQ channels, their importance lies in the ability to modulate neurotransmitter release and somatic excitation in the nervous system. Robust production of PIP2 via hydrolysis agonizes four receptors in the sympathetic neurons of the superior cervical ganglion (e.g., M1, AT1, B2, and P2Y). Modulation of this system occurs via competitive or allosteric regulation of the membrane transport protein affinities for PIP2 molecules [1,5,6,7,8,9,10,11,16,17,18,19,20,21].
With this array of physiological properties found in KCNQ channels, there has been a growth in the literature on KCNQ channel property modifications for therapeutic treatment modalities, as well as the role of these channels in various pathological processes. Specifically, the alteration or loss of function (LOF) by these KCNQ (i.e., channelopathies) highlight their importance in physiological function in the body.
There are various phenotypic presentations of these channelopathies as most are due to genetic etiology amongst whichever genes are involved and the location of the channels, as depicted in Table 1 [17,18,19,20,21]. The most common genes involved in channelopathies are KCNQ1-5 (without consideration of spliced variants) [14,17,21]. KCNQ1 is most expressed in cardiac and cochlear tissue [14,22]. Specifically, cardiac KCNQ1 LOF mutations are associated with type 1 long-QT syndrome [22,23,24,25,26,27,28,29,30]. Cochlear KCNQ1 pathology involves the autosomal recessive long-QT syndrome (Jervell Lange-Nielsen syndrome), which is associated with potassium channelopathy leading to bilateral sensorineural hearing loss as well as the cardiac arrhythmia [26,27,28,29]. KCNQ2 is most expressed in the fetal cerebellum, hippocampus, and medulla [30,31]. Genetic mutation in KCNQ2 is often associated with benign familial neonatal seizures and early-onset epileptic encephalopathy [9,32,33,34,35,36,37,38,39,40,41,42,43,44,45]. KCNQ3 is also most expressed in the fetal cerebellum, hippocampus, and medulla [9,30]. In addition, KCNQ3 mutations are often associated with channelopathies in conjunction with KCNQ2 [33], but additional literature also supports KCNQ mutations in bipolar disorder [36] and various thyroid disorders [37]. Similar to KCNQ1 expression, KCNQ4 is most expressed in the cochlear hair cells but also in trigeminal ganglia [14]. This plays a key role in maintaining the K+ gradient for channel mechanosensation to carry K+ into hair cells to stimulate auditory sensation [14,38]. KCNQ4 mutations are often associated with auditory hearing loss and have therefore been a key target in developing pharmacotherapeutic options for hearing loss [39,40,41,42,43,44,45,46,47,48,49,50,51,52,53,54]. KCNQ5 is most expressed in neural tissue, including the retinal pigment epithelium [48]. However, the lack of recent literature on the profile of these encoded channel subfamilies suggests that there may be unknown channelopathies related to vision homeostasis [14,48]. The expression of these genes is more often in association with other KCNQ genes than what was separately outlined. In addition to KCNQ5, KCNQ1 and KCNQ4 are also often encoded to channels in the neuronal retina and may also have a degree of contribution to its physiological function [9,14,30]. Despite this, the importance of highlighting single gene encoding remains key to approaching neural pathophysiology [2,8,14]. Given this importance, the aim of this review is to provide an up-to-date understanding of the contemporary work of KCNQ channels in order to provide greater emphasis on KCNQ’s involvement in various pathophysiological processes distributed throughout the human body.

2. Modulation of Synaptic Plasticity by KCNQ Channels

There has been a greater development in the role of KCNQ channels among neuronal networks in the past decade. This has led to its consideration for potential pharmacotherapeutic applications [14]. The ability for neuronal modification, or neural plasticity, is a key area of focus in understanding the foundations of learning and memory functions. Anatomically, the origin of the literature on neural plasticity can be further refined by discussing the concept of synaptic plasticity. This concept focuses on hippocampal formation and two principal cell types: pyramidal neurons and granular cells. Specifically, the pyramidal neurons are composed of diverse branching of dendritic neurons, which are responsible for synaptic communication with other neurons [49,50]. The morphological formation of these neurons within the hippocampus leads to the further subfield classification of pyramidal cells in what is known as Cornu Ammonus (CA), divided into CA1, CA2, and CA3 [49,50]. These regions serve an important role in localizing KCNQ channel function in synaptic plasticity [49,51,52,53,54,55,56,57,58,59,60,61]. Within synaptic plasticity, two major models involved in the application of neural plasticity are long-term potentiation (LTP) and long-term depression (LTD) [59]. These models are activity-dependent, and the literature establishes their role in namesake enhancement or reduction in synaptic efficiency. Historically, LTP was initially found in animal models, which found a sustained enhancement in the hippocampus following high-frequency electrode stimulation. LTD was later recognized after laboratory models found the opposite effect following low-frequency simulations [61,62,63,64,65]. At the cellular level, the literature suggests there are numerous factors that play a role in creating the genres of synaptic efficiency and, ultimately, neural plasticity [63,64,65,66,67,68].
Historically, the literature establishes high concentrations of KCNQ2–5 channels in the perisomatic CA1 hippocampal regions [50]. Within dendritic CA1 regions, the current generated by KCNQ channels may not serve as robust of a role as seen in pyramidal CA1 regions [14,50]. Moreover, it has been seen that modulation of KCNQ currents via linopirdine and XE991 do not create effects on synaptic excitability. Rather, it has been shown that the axonal KCNQ channels create a backpropagation into the dendritic CA1 regions [14,65,66,67,68]. This may suggest that the quantity of KCNQ channels in the dendrites does not play as robust of a role in synaptic excitability as the axonal KCNQ channels do themselves [64,65,66,67,68]. This makes axonal KCNQ channels the greater focus of study.
It was initially found that pharmacologic KCNQ channel inhibition via linopirdine reduced spike frequency adaptation (SFA) in CA1 pyramidal neurons in vitro, but only after the initial spike [49,50]. Following this initial discovery, it was also found that KCNQ channel modulation also plays a role in after hyperpolarization, which ultimately supports the notion that KCNQ channels contribute to AP [66]. In addition, muscarinic channel inhibition (i.e., KCNQ) has been shown to stimulate an array of homeostatic neuroplastic changes in synaptic efficiency. This array, in combination with KCNQ’s contribution to AP, may suggest that this array occurs at different time points, which allows for understanding that a temporal process of these neuroplastic changes occurs rather than a synced process in the hippocampus [67]. If the behavior of KCNQ channels occurs in a temporal process, this can make way for a greater understanding of the role of KCNQ via LTP and, therefore, memory development.
LTP genre can be categorized as either dependent or independent of N-methyl-D-aspartate (NMDA) receptors. Within the NMDA receptor-dependent form of LTP, it is suggested that KCNQ inhibition via XE991 stimulates the opening of NMDA receptors mediated channels during LTP by stimulating the depolarization after AP firing when performed via theta-burst stimulation [65]. This behavior may suggest that XE991 inhibition could serve a pharmacotherapeutic role in improving memory. However, the literature regarding the modulation of KCNQ channels regarding memory and LTP is still not completely understood [62,63,64,65,66,67,68]. With regard to LTP in the presence of acute stress, it is well understood that stress impairs spatial memory retrieval. Flupirtine-induced activation of KCNQ channels in the CA1 region is found to have a neuroprotective effect on spatial memory retrieval in the case of acute stress. The mechanism behind these protective effects is suggested to be through the Akt/GSK-3β and Erk1/2 signaling pathways, and animal models have shown flupirtine treatments resulted in decreased expression of apoptosis factors (i.e., Bax) and upregulation of hippocampal p-Erk1/2 [66,67,68]. Likewise, literature establishes beneficial effects on memory via KCNQ pharmacological inhibition as well. In addition to the aforementioned effects on LTP by XE991, this inhibitor has also improved cognitive impairment secondary to acetylcholine (ACh) depletion animal model induced by the neuroexcitatory kainic acid [64,65,66,67,68]. The discovery of this behavior has led to a growing body of literature on potential therapeutic applications in Alzheimer’s Disease due to its nature as a cholinergic deficiency-related cognitive impairment [66,67]. In addition, the inhibition of KCNQ channels via linopirdine is also well-established in enhancing cognition via increased ACh release [63,65,66,67,68,69,70,71,72,73,74,75,76,77].
In contrast, while LTP typically occurs after a brief high-intensity stimulation of a postsynaptic neuron, LTD can be caused by prolonged low-intensity stimulation or simulation that occurs after the firing of an AP [69]. This leads to insufficient depolarization due to this lower level of stimulation. This does not generate a removal of the magnesium blockage of the NMDA receptor [78,79,80,81,82]. However, there is evidence that this stimulation is enough to open some NMDA receptors to allow for calcium ions into the cell. These cellular calcium levels are thought to activate a cellular cascade to remove α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid (AMPA) receptors [77,78,79,80,81]. This reduces the postsynaptic glutamate receptor density, which decreases synapse efficiency and, therefore, memory and learning development. Despite the developments in literature dedicated to LTD, there is little literature on the effects that KCNQ channels have on this mechanism compared to LTP.
Other than the pharmacological inhibition of KCNQ channels, the inhibition by genetic proxy also serves a role in neural plasticity with regard to cognition. Animal models have shown epileptic seizures in addition to cognitive spatial memory impairment in cases of mutant or LOF genes that encode for KCNQ2 [82,83,84,85,86,87,88,89]. These effects brought upon by genetic inhibition challenge the protective and cognitive improvements seen in pharmacologic forms of KCNQ inhibition. This is where the literature ought to focus in order to determine if secondary factors that are not understood in these animal models also contribute to the memory impairment (i.e., hippocampal morphology) [90,91,92,93,94,95]. The epileptic phenotype, in conjunction with the cognitive impairment of these genetic models, may suggest additional psychomotor exploration [95,96,97,98,99,100,101,102,103,104].
Clinically, neuroplastic changes to the central nervous system are well documented in the case of chronic pain [105,106,107,108]. Moreover, there is a large number of literatures that focuses on the ability to enhance positive neuroplasticity as a clinical application for chronic pain treatment. Potassium channels can be regulated for the function of membrane excitability. The dorsal horn neurons and the DRG sensory neurons express the neuronal Kv7 channels, which are formed by Kv7.2, Kv7.3, and Kv7.5 subunits [80]. Depolarization of the resting membrane potential and increasing firing of nociceptive neurons occurs when Kv7 channels have either been blocked or have experienced a LOF. As such, a potential treatment for chronic pain can be increasing the function of such Kv7 channels in nociceptors. Retigabine, a Kv7 opener, was initially investigated as a potential seizure treatment but has been discontinued for concerns of toxicity and skin/retinal discoloration [109,110,111]. Recently, SCR2682 has been shown to decrease DRG neuron excitability in vitro and strongly activate Kv7 currents in neurons. In persistent, inflammatory, neuropathic pain models, SCR2862 activation of Kv7 currents reduced the thermal hyperalgesia and mechanical allodynia [112].
In addition to Retigabine and XE991, the anti-inflammatory painkillers celecoxib and diclofenac could potentially impact Kv7 channels, underlining the importance of Kv7 as a potential analgesic target [84]. Diclofenac can directly depress spinal nociceptive transmission and spinal reflexes. The literature by Baz et al. has emphasized that both diclofenac and flupirtine spinal effects are mediated through KCNQ channels [111,112,113,114,115]. Another study investigated the anti-nociceptive effects on chronic and inflammatory pain of the novel opener QO58-lysine on KCNQ channels, finding that there is a pharmacological effect of QO58-lysine effect on pain [85,86]. Inflammatory pain was reversed by administration of QO58-lysine without toxicity as well [85,86,87,88,89].

3. KCNQ Channels in Dopaminergic Motor Disorders

KCNQ channels have become important targets in disorders regarding dopaminergic pathways. Channels of this group in the ventral tegmental area (VTA) of the mesolimbic dopaminergic tract have been explored as targets of pain attenuation [80]. While this avenue of study has proven to be a promising utilization of these K+ channels in dopaminergic pathways, they are perhaps better associated with the treatment of motor disorders involving the neurotransmitter dopamine. Parkinson’s disease (PD) is a neurodegenerative disorder commonly associated with the loss of dopaminergic neurons in the substantia nigra pars compacta (SNc) [90,91,92,93,94,95,96,97,98,99,100,101,102,103]. As a result, the loss of nigrostriatal dopamine is associated with the motor symptoms classically observed in PD, such as muscular rigidity, bradykinesia, and tremor [84,85,86,87,88,89,90,91,92,93,94,95]. Recently, it has been suggested that the modulation of certain K+ channels is able to elicit protective effects upon nigrostriatal dopaminergic neurons, attenuating the motor symptoms observed in PD. Among the K+ channels capable of attenuating the motor-related symptoms of PD is the KCNQ family of K+ channels. KCNQ2 and KCNQ4 specifically have been shown to exist in high concentrations in the SNc compared to other members of the KCNQ family [95].
Activators of the KCNQ channels in the substantia nigra, such as retigabine (a KCNQ2-5 opener), can induce an inhibitory control of nigrostriatal neurons through increased K+ conductance [96,97]. This effect is able to attenuate the activity of dopaminergic neurons, diminishing the amount of dopamine released in the SNc. Conversely, the blockade of the KCNQ channels via XE991 had the opposite effect. Through the inhibition of K+ conductance in the dopaminergic neurons of the SNc, XE991 can induce greater excitability in these neurons [80]. It has therefore been hypothesized that XE991 may be a potential pharmacotherapeutic option in the treatment of PD. By targeting KCNQ channels in the SNc, increasing the generation of action potentials in existing dopaminergic neurons may potentially alleviate the symptoms caused by a loss of dopaminergic neurons, as seen in PD. A recent study conducted by Chen et al. attempted to determine the efficacy of XE991 in attenuating the motor symptoms associated with PD in vivo. Catalepsy was induced in rats treated with intraperitoneal injections of haloperidol to create a PD-like state. The study found that administration of XE991 into the SNc of the rats was able to alleviate the symptoms of catalepsy induced by systemic administration of haloperidol, providing an in-vivo example of the treatment of motor symptoms in a PD-like state via targeting of KCNQ channels [98].
In addition to increasing the excitability of dopaminergic neurons in the SNc, XE991 has been shown to provide neuroprotective effects as well by preventing dopaminergic neuronal death. The potent neurotoxin, 6-hydroxydopamine (6-OHDA), has been shown to replicate an irreversible state similar to PD by the destruction of dopaminergic neurons in the SNc [98,99]. A recent study conducted by Liu et al. attempted to observe the neuroprotective effects of XE991 on 6-OHDA treated rats [80]. During the study, it was found that rats with 6-OHDA injections in the medial forebrain bundle (MFB) exhibited significantly lower levels of tyrosine hydroxylase positive neurons in the SNc, compared to rats treated with a sham injection in the MFB. Additionally, another group of rats was treated with 6-OHDA as well as XE991 to observe the potential neuroprotective effects of blocking KCNQ channels. It was found via immunofluorescence that the survival ratio of tyrosine hydroxylase positive neurons was significantly greater in the group treated with XE991 and 6-OHDA (54.43%) than in the group treated only with 6-OHDA (16.63%), leading to the conclusion that XE991 was able to prevent attenuation of nigrostriatal dopamine in a PD-like state [80].
The neuroprotective effects associated with the blockade of KCNQ channels by XE991 may potentially be associated with an attenuation of motor symptoms associated with PD. In an open-field test, 6-OHDA injected mice treated with XE991 exhibited significantly greater walking speeds than mice treated only with 6-OHDA, displaying the ability of XE991 to counteract bradykinetic symptoms. A balance beam test designed to assess balance and coordination was also conducted, with higher scores representing a greater ability to traverse a balance beam. It was once again found that 6-OHDA injected rats treated with XE991 displayed significantly greater scores than their counterparts only injected with 6-OHDA. Finally, the study conducted a grip test in which rats who were able to hold onto a wire for a longer period of time exhibited greater muscular rigidity, a classic symptom of PD. The study found a significant decrease in grip time (and therefore, muscular rigidity) in rats injected with 6-OHDA treated with XE991 compared to rats only injected with 6-OHDA. To further explore the protective properties of XE991, retigabine was applied to another group injected with 6-OHDA and XE991. In this group, slower walking speeds, decreased balance beam scores, and increased grip time were noted compared to the group treated with XE991. The results of this analysis further support the neuroprotective capabilities of XE991, as activation of KCNQ channels via retigabine reverses the benefits of XE991 [80].
Although retigabine can reverse the XE991-related attenuation of motor symptoms in 6-OHDA treated rats, it has been hypothesized to diminish hyperkinetic motor symptoms caused by long-term treatment of PD. Considering that hypoactivity of dopaminergic neurons induced by PD-like states has been treated by the blockade of KCNQ channels via XE991, it would seem logical that activation of KCNQ channels via retigabine and other channel openers would attenuate symptoms due to dopaminergic hyperactivity seen in the treatment of PD. Currently, one of the primary treatment methods for PD includes the compound L-DOPA. Within 10 years of L-DOPA administration, many PD patients receiving this drug tend to become afflicted by hyperkinetic symptoms such as chorea, myoclonus, or dystonia [99,100,101]. In rat models of L-DOPA-induced dyskinesia (LID), it was found that retigabine significantly reduced the frequency of abnormal involuntary movements (AIM). However, it is important to note that retigabine was unable to completely prevent the occurrence of LID or delay the onset, even with chronic treatment. A limitation of retigabine as a pharmacotherapeutic option for the treatment of LID is its potential to induce sedation [102,103,104,105]. This effect, however, was reduced over the course of chronic treatment with retigabine. Similarly, the KCNQ2/3 opener, N-(6-chloro-pyridin-3-yl)-3,4-difluoro-benzamide (ICA 27243), was also found to significantly diminish the frequency of AIMs seen in rat models of PD [104,105]. It is hypothesized that both retigabine and ICA 27243 reduce AIM frequency by acting upon KCNQ2/3 in striatal projection neurons, whose increase in activity is associated with the symptoms of LTD [104,105].
As recent studies have shown, the modulation of dopaminergic activity via KCNQ channels has the potential to treat several motor disorders associated with both LID and Parkinsonian states. Due to the relative lack of side effects associated with drugs that modulate the activity of KCNQ channels, these compounds appear to be promising new avenues of pharmacotherapy in the management of PD. Further understanding of the therapeutic potential of these drugs is essential due to the side effects associated with current PD treatments, such as L-DOPA. A future direction should attempt to discern the existence of unknown adverse effects through continued trials of these drugs in PD murine models, as these drugs have only minimally been utilized in studies regarding the attenuation of motor symptoms. Additionally, future research into the ability of KCNQ modulators to treat other disorders related to dopaminergic pathways is another avenue that should be explored as well. Past studies have speculated that KCNQ openers, such as retigabine and ICA 27243, may be able to reduce the excitability of dopaminergic neurons in other tracts, such as the mesolimbic pathway, where overactivity results in schizophrenic symptoms [105,106,107,108,109,110,111,112,113]. This could be a potentially innovative tactic in clinical medicine as well, especially in the case of dementia-related psychosis, which is common onset among those with neurodegenerative disease. Moreover, recent clinical trials on the oral use of primavanserin, an atypical antipsychotic which has antagonist/inverse agonist effects on 5HT2A receptors, have been currently conducted. Within the most recent trial, a side effect of primavanserin use as a treatment for PD was QT prolongation [114]. This could be the particular area in which KCNQ channel modulation (e.g., retigabine), could aid in limiting these side effects. However, no current studies have yet to be carried out on this potential dual medication use. The KCNQ channel opening and closing mechanisms by selected agents are listed in Table 2. Overall, the ability of KCNQ channels to modulate a wide variety of motor-related dopaminergic pathologies is a promising field of study and may provide insight into the treatment of a wide variety of dopaminergic disorders in the future.

4. Discussion

The aim of this review was to encompass the updated literature and scope of KCNQ channel behavior in neural mechanosensation pathways. This aim was achieved by establishing the concept of VGICs, and the specific structure and physiology of the KCNQ family and related to both motor and sensory circuits. The novelty behind this updated review is that it encompasses relevant channelopathies as well as conceptual pharmacotherapeutic options via modulation of the KCNQ channel, as depicted in Figure 2. However, the future direction of KCNQ neural mechanosensation literature ought to focus more on the clinical and translational use of these pharmacotherapeutic options, as most of the literature used to support this review was based on animal models alone. Despite this, there is strong evidence that could support the initiation of future clinical trials, especially for auditory pathologies as well as in PD [14,80,94,95,96,97,98,99,100,101,102,103,104,105,115,116,117,118,119,120,121,122], and developments for understanding the effects that antimicrobial agents have on these potassium ion channels [123,124,125,126,127,128,129,130,131,132,133,134,135,136,137,138]. Additionally, the scope of this review was on neural modulation, yet there is a growing amount of literature on the behavior of KCNQ channels in cardiac modulations as well, especially in the cellular processes involved in arrhythmia [26,27,28,29,139,140,141,142,143,144,145,146,147,148,149].

5. Conclusions

Overall, the aim of this review was to encompass the current literature and scope of KCNQ channels. KCNQ channels remain an imperative area of pathophysiological study. Some of the most pertinent roles of these channels include their mechanosensation behavior in dopaminergic motor disorders, pain sensation, neural plasticity, and cranial sensory transduction (e.g., hearing, olfaction, vision). Future directions ought to include exploring these treatment modalities in clinical trials as well as understanding cellular processes with non-neural disorders just as well.

Author Contributions

Conceptualization, X.-P.C. and S.P.S.; software, M.M. and S.P.S.; resources, X.-P.C.; writing—original draft preparation, S.P.S., M.W., M.M. and X.-P.C.; writing—review and editing, S.P.S., M.W., M.M. and X.-P.C.; visualization, M.M. and S.P.S.; supervision, X.-P.C.; project administration, X.-P.C.; funding acquisition, X.-P.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by American Heart Association (grant number: 19AIREA34470007).

Institutional Review Board Statement

Not applicable.

Informed Consent 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., M.W. and M.M.’s professional studies. S.S. and M.W. are recipients of 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

All authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Abbreviations

Achacetylcholine
AIMabnormal involuntary movements
AMPAα-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid (AMPA)
APaction potential
CAcornu ammonus
DRGdorsal root ganglion
ICA 27243N-(6-chloro-pyridin-3-yl)-3,4-difluoro-benzamide
KCNQvoltage-gated potassium channel
Kvvoltage-gated potassium channel
LIDL-DOPA induced dyskinesia
LOFloss of function
LTDlong-term depression
LTPlong-term potentiation
mAChRmuscarinic acetylcholine receptors
MFBmedial forebrain bundle
NMDAN-methyl-D-aspartate
6-OHDA6-hydroxydopamine
PDParkinson’s disease
PIP2phosphatidylinositol 4,5-bisphosphate
SFAspike frequency adaptation
SNcsubstantia nigra pars compacta
VGICsvoltage-gated potassium channels
VGKCvoltage-gated potassium channel
VTAventral tegmental area

References

  1. Camerino, D.C.; Desaphy, J.F.; Tricarico, D.; Pierno, S.; Liantonio, A. Therapeutic approaches to ion channel diseases. Adv. Genet. 2008, 64, 81–145. [Google Scholar]
  2. Abbott, G.W. KCNQs: Ligand- and voltage-gated potassium channels. Front. Physiol. 2020, 11, 583. [Google Scholar] [CrossRef]
  3. Kefauver, J.M.; Ward, A.B.; Patapoutian, A. Discoveries in structure and physiology of mechanically activated ion channels. Nature 2020, 587, 567–576. [Google Scholar] [CrossRef]
  4. Häfner, S.; Sandoz, G. Photopharmacological approaches for dissecting potassium channel physiology. Curr. Opin. Pharmacol. 2022, 63, 102178. [Google Scholar] [CrossRef]
  5. Mandal, K. Review of PIP2 in cellular signaling, functions and diseases. Int. J. Mol. Sci. 2020, 21, 8342. [Google Scholar] [CrossRef]
  6. Harraz, O.F. PIP2: A critical regulator of vascular ion channels hiding in plain sight. Proc. Natl. Acad. Sci. USA 2020, 117, 20378–20389. [Google Scholar] [CrossRef]
  7. Krajnik, A.; Brazzo, J.A.; Vaidyanathan, K.; Das, T.; Redondo-Muñoz, J.; Bae, Y. Phosphoinositide signaling and mechanotransduction in cardiovascular biology and disease. Front. Cell Dev. Biol. 2020, 8, 595849. [Google Scholar] [CrossRef]
  8. McLean, M.A.; Stephen, A.G.; Sligar, S.G. PIP2 influences the conformational dynamics of membrane-bound KRAS4b. Biochemistry 2019, 58, 3537–3545. [Google Scholar] [CrossRef]
  9. Jespersen, T.; Grunnet, M.; Olesen, S.P. The KCNQ1 potassium channel: From gene to physiological function. Physiology 2005, 20, 408–416. [Google Scholar] [CrossRef] [Green Version]
  10. Liu, W.X.; Deng, E.Z.; Chen, W.; Lin, H. Identifying the subfamilies of voltage-gated potassium channels using feature selection technique. Int. J. Mol. Sci. 2014, 15, 12940–12951. [Google Scholar] [CrossRef] [Green Version]
  11. Ranjan, R.; Logette, E.; Marani, M.; Herzog, M.; Tâche, V.; Scantamburlo, E.; Buchillier, V.; Markram, H. A Kinetic map of the homomeric voltage-gated potassium channel (Kv) family. Front. Cell. Neurosci. 2019, 13, 358. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Brown, D.A.; Adams, P.R. Muscarinic suppression of a novel voltage-sensitive K+ current in a vertebrate neurone. Nature 1980, 283, 673–676. [Google Scholar] [CrossRef] [PubMed]
  13. Delmas, P.; Brown, D.A. Pathways modulating neural KCNQ/M (Kv7) potassium channels. Nat. Rev. Neurosci. 2005, 6, 850–862. [Google Scholar] [CrossRef] [PubMed]
  14. Wang, J.J.; Li, Y. KCNQ potassium channels in sensory system and neural circuits. Acta Pharmacol. Sin. 2016, 37, 25–33. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Eren-Koçak, E.; Dalkara, T. Ion channel dysfunction and neuroinflammation in migraine and depression. Front. Pharmacol. 2021, 12, 777607. [Google Scholar] [CrossRef] [PubMed]
  16. Sacco, T.; Tempia, F. A-type potassium currents active at subthreshold potentials in mouse cerebellar Purkinje cells. J. Physiol. 2002, 543, 505–520. [Google Scholar] [CrossRef]
  17. Brown, D.A.; Passmore, G.M. Neural KCNQ (Kv7) channels. Br. J. Pharmacol. 2009, 156, 1185–1195. [Google Scholar] [CrossRef]
  18. Jentsch, T.J. Neuronal KCNQ potassium channels: Physiology and role in disease. Nat. Rev. Neurosci. 2000, 1, 21–30. [Google Scholar] [CrossRef]
  19. Lehmann-Horn, F.; Jurkat-Rott, K. Voltage-gated ion channels and hereditary disease. Physiol. Rev. 1999, 79, 1317–1372. [Google Scholar] [CrossRef]
  20. Jurkat-Rott, K.; Lehmann-Horn, F. Human muscle voltage-gated ion channels and hereditary disease. Curr. Opin. Pharmacol. 2001, 1, 280–287. [Google Scholar] [CrossRef]
  21. Felix, R. Channelopathies: Ion channel defects linked to heritable clinical disorders. J. Med. Genet. 2000, 37, 729–740. [Google Scholar] [CrossRef] [PubMed]
  22. Robbins, J. KCNQ potassium channels: Physiology, pathophysiology, and pharmacology. Pharmacol. Ther. 2001, 90, 1–19. [Google Scholar] [CrossRef]
  23. Wang, Z.; Wang, L.; Liu, W.; Hu, D.; Gao, Y.; Ge, Q.; Liu, X.; Li, L.; Wang, Y.; Wang, S.; et al. Pathogenic mechanism and gene correction for LQTS-causing double mutations in KCNQ1 using a pluripotent stem cell model. Stem Cell Res. 2019, 38, 101483. [Google Scholar] [CrossRef] [PubMed]
  24. Wei, H.; Wu, J.; Liu, Z. Studying KCNQ1 mutation and drug response in type 1 long QT syndrome using patient-specific induced pluripotent stem cell-derived cardiomyocytes. Methods Mol. Biol. 2018, 1684, 7–28. [Google Scholar]
  25. Ma, D.; Wei, H.; Lu, J.; Huang, D.; Liu, Z.; Loh, L.J.; Islam, O.; Liew, R.; Shim, W.; Cook, S.A. Characterization of a novel KCNQ1 mutation for type 1 long QT syndrome and assessment of the therapeutic potential of a novel IKs activator using patient-specific induced pluripotent stem cell-derived cardiomyocytes. Stem Cell Res. Ther. 2015, 6, 39. [Google Scholar] [CrossRef]
  26. Wuriyanghai, Y.; Makiyama, T.; Sasaki, K.; Kamakura, T.; Yamamoto, Y.; Hayano, M.; Harita, T.; Nishiuchi, S.; Chen, J.; Kohjitani, H.; et al. Complex aberrant splicing in the induced pluripotent stem cell-derived cardiomyocytes from a patient with long QT syndrome carrying KCNQ1-A344Aspl mutation. Heart Rhythm 2018, 15, 1566–1574. [Google Scholar] [CrossRef]
  27. García Gozalo, M.; Bermejo Arnedo, I.; De Vera McMullan, P. KCNQ1 gene mutation and epilepsy in patient with long QT syndrome. Med. Clin. 2021, 157, 456–457. [Google Scholar] [CrossRef]
  28. Marstrand, P.; Almatlouh, K.; Kanters, J.K.; Graff, C.; Christensen, A.H.; Bundgaard, H.; Theilade, J. Effect of moderate potassium-elevating treatment in long QT syndrome: The TriQarr potassium study. Open Heart 2021, 8, e001670. [Google Scholar] [CrossRef]
  29. Zhang, R.; Ding, C.; Wang, H. Treatment on arrhythmia electric storm in a Jervell and Lange-Nielsen syndrome patient by ablation of the triggering premature ventricular contraction: A case report. Ann. Palliat. Med. 2021, 10, 4938–4943. [Google Scholar] [CrossRef]
  30. Giudicessi, J.R.; Ackerman, M.J. Prevalence and potential genetic determinants of sensorineural deafness in KCNQ1 homozygosity and compound heterozygosity. Circ. Cardiovasc. Genet. 2013, 6, 193–200. [Google Scholar] [CrossRef] [Green Version]
  31. Qiu, Y.; Chen, S.; Wu, X.; Zhang, W.J.; Xie, W.; Jin, Y.; Xie, L.; Xu, K.; Bai, X.; Zhang, H.M.; et al. Jervell and Lange-Nielsen syndrome due to a novel compound heterozygous KCNQ1 mutation in a Chinese family. Neural Plast. 2020, 2020, 3569359. [Google Scholar] [CrossRef] [PubMed]
  32. Vyas, B.; Puri, R.D.; Namboodiri, N.; Nair, M.; Sharma, D.; Movva, S.; Saxena, R.; Bohora, S.; Aggarwal, N.; Vora, A.; et al. KCNQ1 mutations associated with Jervell and Lange-Nielsen syndrome and autosomal recessive Romano-Ward syndrome in India-expanding the spectrum of long QT syndrome type 1. Am. J. Med. Genet. 2016, 170, 1510–1519. [Google Scholar] [CrossRef] [PubMed]
  33. Yang, Q.; Tan, Q.Q.; Lan, C.J.; Lv, B.Z.; Zhou, G.M.; Zhong, W.Q.; Gu, Z.M.; Mao, Y.M.; Liao, X. The changes of KCNQ5 expression and potassium microenvironment in the retina of myopic guinea pigs. Front. Physiol. 2021, 12, 790580. [Google Scholar] [CrossRef] [PubMed]
  34. Mönnig, G.; Schulze-Bahr, E.; Wedekind, H.; Eckardt, L.; Kirchhof, P.; Funke, H.; Kotthoff, S.; Vogt, J.; Assmann, G.; Breithardt, G.; et al. Clinical aspects and molecular genetics of the Jervell- and Lange-Nielsen Syndrome. Z. Kardiol. 2002, 91, 380–388. [Google Scholar] [CrossRef] [PubMed]
  35. Kanaumi, T.; Takashima, S.; Iwasaki, H.; Itoh, M.; Mitsudome, A.; Hirose, S. Developmental changes in KCNQ2 and KCNQ3 expression in human brain: Possible contribution to the age-dependent etiology of benign familial neonatal convulsions. Brain Dev. 2008, 30, 362–369. [Google Scholar] [CrossRef] [PubMed]
  36. Devaux, J.J.; Kleopa, K.A.; Cooper, E.C.; Scherer, S.S. KCNQ2 is a nodal K+ channel. J. Neurosci. 2004, 24, 1236–1244. [Google Scholar] [CrossRef] [Green Version]
  37. Mary, L.; Nourisson, E.; Feger, C.; Laugel, V.; Chaigne, D.; Keren, B.; Afenjar, A.; Billette, T.; Trost, D.; Cieuta-Walti, C.; et al. Pathogenic variants in KCNQ2 cause intellectual deficiency without epilepsy: Broadening the phenotypic spectrum of a potassium channelopathy. Am. J. Med. Genet. A 2021, 185, 1803–1815. [Google Scholar] [CrossRef]
  38. Vanoye, C.G.; Desai, R.R.; Ji, Z.; Adusumilli, S.; Jairam, N.; Ghabra, N.; Joshi, N.; Fitch, E.; Helbig, K.L.; McKnight, D.; et al. High-throughput evaluation of epilepsy-associated KCNQ2 variants reveals functional and pharmacological heterogeneity. JCI Insight 2022, 7, e156314. [Google Scholar] [CrossRef]
  39. Hu, C.; Liu, D.; Luo, T.; Wang, Y.; Liu, Z. Phenotypic spectrum and long-term outcome of children with genetic early-infantile-onset developmental and epileptic encephalopathy. Epileptic Disord. 2022. [Google Scholar] [CrossRef]
  40. Kim, K.W.; Kim, K.; Kim, H.J.; Kim, B.I.; Baek, M.; Suh, B.C. Posttranscriptional modulation of KCNQ2 gene expression by the miR-106b microRNA family. Proc. Natl. Acad. Sci. USA 2021, 118, e2110200118. [Google Scholar] [CrossRef]
  41. Monni, L.; Kraus, L.; Dipper-Wawra, M.; Soares-Da-Silva, P.; Maier, N.; Schmitz, D.; Holtkamp, M.; Fidzinski, P. In vitro and in vivo anti-epileptic efficacy of eslicarbazepine acetate in a mouse model of KCNQ2-related self-limited epilepsy. Br. J. Pharmacol. 2022, 179, 84–102. [Google Scholar] [CrossRef] [PubMed]
  42. Nissenkorn, A.; Kornilov, P.; Peretz, A.; Blumkin, L.; Heimer, G.; Ben-Zeev, B.; Attali, B. Personalized treatment with retigabine for pharmacoresistant epilepsy arising from a pathogenic variant in the KCNQ2 selectivity filter. Epileptic Disord. 2021, 23, 695–705. [Google Scholar] [CrossRef] [PubMed]
  43. Lee, I.C.; Chang, T.M.; Liang, J.S.; Li, S.Y. KCNQ2 mutations in childhood nonlesional epilepsy: Variable phenotypes and a novel mutation in a case series. Mol. Genet. Genom. Med. 2019, 7, e00816. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Milh, M.; Lacoste, C.; Cacciagli, P.; Abidi, A.; Sutera-Sardo, J.; Tzelepis, I.; Colin, E.; Badens, C.; Afenjar, A.; Coeslier, A.D.; et al. Variable clinical expression in patients with mosaicism for KCNQ2 mutations. Am. J. Med. Genet. A 2015, 167, 2314–2318. [Google Scholar] [CrossRef] [PubMed]
  45. Lazo, P.A.; García, J.L.; Gómez-Puertas, P.; Marcos-Alcalde, Í.; Arjona, C.; Villarroel, A.; González-Sarmiento, R.; Fons, C. Novel dominant KCNQ2 exon 7 partial in-frame duplication in a complex epileptic and neurodevelopmental delay syndrome. Int. J. Mol. Sci. 2020, 21, 4447. [Google Scholar] [CrossRef]
  46. Kaminsky, Z.; Jones, I.; Verma, R.; Saleh, L.; Trivedi, H.; Guintivano, J.; Akman, R.; Zandi, P.; Lee, R.S.; Potash, J.B. DNA methylation and expression of KCNQ3 in bipolar disorder. Bipolar Disord. 2015, 17, 150–159. [Google Scholar] [CrossRef]
  47. Mittal, K.; Rafiq, M.A.; Rafiullah, R.; Harripaul, R.; Ali, H.; Ayaz, M.; Aslam, M.; Naeem, F.; Amin-Ud-Din, M.; Waqas, A.; et al. Mutations in the genes for thyroglobulin and thyroid peroxidase cause thyroid dyshormonogenesis and autosomal-recessive intellectual disability. J. Hum. Genet. 2016, 61, 867–872. [Google Scholar] [CrossRef]
  48. Rim, J.H.; Choi, J.Y.; Jung, J.; Gee, H.Y. Activation of KCNQ4 as a therapeutic strategy to treat hearing loss. Int. J. Mol. Sci. 2021, 22, 2510. [Google Scholar] [CrossRef]
  49. Lee, S.Y.; Choi, H.B.; Park, M.; Choi, I.S.; An, J.; Kim, A.; Kim, E.; Kim, N.; Han, J.H.; Kim, M.Y.; et al. Novel KCNQ4 variants in different functional domains confer genotype- and mechanism-based therapeutics in patients with nonsyndromic hearing loss. Exp. Mol. Med. 2021, 53, 1192–1204. [Google Scholar] [CrossRef]
  50. Thorpe, R.K.; Walls, W.D.; Corrigan, R.; Schaefer, A.; Wang, K.; Huygen, P.; Casavant, T.L.; Smith, R.J.H. AudioGene: Refining the natural history of KCNQ4, GSDME, WFS1, and COCH-associated hearing loss. Hum. Genet. 2022, 141, 877–887. [Google Scholar] [CrossRef]
  51. Yen, T.T.; Chen, I.C.; Hua, M.W.; Wei, C.Y.; Shih, K.H.; Li, J.L.; Lin, C.H.; Hsiao, T.H.; Chen, Y.M.; Jiang, R.S. A KCNQ4 c.546C>G Genetic variant associated with late onset non-syndromic hearing loss in a Taiwanese population. Genes 2021, 12, 1711. [Google Scholar] [CrossRef] [PubMed]
  52. Kojima, T.; Wasano, K.; Takahashi, S.; Homma, K. Cell death-inducing cytotoxicity in truncated KCNQ4 variants associated with DFNA2 hearing loss. Dis. Model Mech. 2021, 14, dmm049015. [Google Scholar] [CrossRef] [PubMed]
  53. Peixoto-Pinheiro, B.; Vona, B.; Löwenheim, H.; Rüttiger, L.; Knipper, M.; Adel, Y. Age-related hearing loss pertaining to potassium ion channels in the cochlea and auditory pathway. Dis. Model Mech. 2021, 473, 823–840. [Google Scholar] [CrossRef] [PubMed]
  54. Borgini, M.; Mondal, P.; Liu, R.; Wipf, P. Chemical modulation of Kv7 potassium channels. RSC Med. Chem. 2021, 12, 483–537. [Google Scholar] [CrossRef] [PubMed]
  55. Xiong, Q.; Sun, H.; Li, M. Zinc pyrithione-mediated activation of voltage-gated KCNQ potassium channels rescues epileptogenic mutants. Nat. Chem. Biol. 2007, 3, 287–296. [Google Scholar] [CrossRef]
  56. Tompson, D.J.; Buraglio, M.; Andrews, S.M.; Wheless, J.W. Adolescent clinical development of ezogabine/retigabine as adjunctive therapy for partial-onset seizures: Pharmacokinetics and tolerability. J. Pediatr. Pharmacol. Ther. 2016, 21, 404–412. [Google Scholar] [CrossRef] [Green Version]
  57. Gunthorpe, M.J.; Large, C.H.; Sankar, R. The mechanism of action of retigabine (ezogabine), a first-in-class K+ channel opener for the treatment of epilepsy. Epilepsia 2012, 53, 412–424. [Google Scholar] [CrossRef]
  58. Bayasgalan, T.; Stupniki, S.; Kovács, A.; Csemer, A.; Szentesi, P.; Pocsai, K.; Dionisio, L.; Spitzmaul, G.; Pál, B. Alteration of mesopontine cholinergic function by the lack of KCNQ4 subunit. Front. Cell. Neurosci. 2021, 15, 707789. [Google Scholar] [CrossRef]
  59. Niu, X.; Yu, K.; He, B. Transcranial focused ultrasound induces sustained synaptic plasticity in rat hippocampus. Brain Stimul. 2022, 15, 352–359. [Google Scholar] [CrossRef]
  60. Caragea, V.M.; Manahan-Vaughan, D. Bidirectional regulation of hippocampal synaptic plasticity and modulation of cumulative spatial memory by dopamine D2-like receptors. Front. Behav. Neurosci. 2022, 15, 803574. [Google Scholar] [CrossRef]
  61. Sahu, G.; Turner, R.W. The molecular basis for the calcium-dependent slow afterhyperpolarization in CA1 hippocampal pyramidal neurons. Front. Physiol. 2021, 12, 759707. [Google Scholar] [CrossRef]
  62. Bentzen, B.H.; Schmitt, N.; Calloe, K.; Dalby, B.W.; Grunnet, M.; Olesen, S.P. The acrylamide (S)-1 differentially affects Kv7 (KCNQ) potassium channels. Neuropharmacology 2006, 51, 1068–1077. [Google Scholar] [CrossRef]
  63. Blom, S.M.; Schmitt, N.; Jensen, H.S. The acrylamide (S)-2 as a positive and negative modulator of Kv7 channels expressed in Xenopus laevis oocytes. PLoS ONE 2009, 4, e8251. [Google Scholar] [CrossRef]
  64. Zhang, X.; An, H.; Li, J.; Zhang, Y.; Liu, Y.; Jia, Z.; Zhang, W.; Chu, L.; Zhang, H. Selective activation of vascular Kv 7.4/Kv 7.5 K+ channels by fasudil contributes to its vasorelaxant effect. Br. J. Pharmacol. 2016, 173, 3480–3491. [Google Scholar] [CrossRef] [Green Version]
  65. Zhang, X.; Yang, D.; Hughes, B.A. KCNQ5/K(v)7.5 potassium channel expression and subcellular localization in primate retinal pigment epithelium and neural retina. Am. J. Physiol. Cell Physiol. 2011, 301, C1017–C1026. [Google Scholar] [CrossRef] [Green Version]
  66. Fogwe, L.A.; Reddy, V.; Mesfin, F.B. Neuroanatomy, Hippocampus; StatPearls Publishing: Treasure Island, FL, USA, 2021. Available online: https://www.ncbi.nlm.nih.gov/books/NBK482171 (accessed on 15 January 2020).
  67. Hu, H.; Vervaeke, K.; Storm, J.F. M-channels (Kv7/KCNQ channels) that regulate synaptic integration, excitability, and spike pattern of CA1 pyramidal cells are located in the perisomatic region. J. Neurosci. 2007, 27, 1853–1867. [Google Scholar] [CrossRef] [Green Version]
  68. Cooper, E.C.; Harrington, E.; Jan, Y.N.; Jan, L.Y. M channel KCNQ2 subunits are localized to key sites for control of neuronal network oscillations and synchronization in mouse brain. J. Neurosci. 2001, 21, 9529–9540. [Google Scholar] [CrossRef] [Green Version]
  69. Tzingounis, A.V.; Heidenreich, M.; Kharkovets, T.; Spitzmaul, G.; Jensen, H.S.; Nicoll, R.A.; Jentsch, T.J. The KCNQ5 potassium channel mediates a component of the afterhyperpolarization current in mouse hippocampus. Proc. Natl. Acad. Sci. USA 2010, 107, 10232–10237. [Google Scholar] [CrossRef] [Green Version]
  70. Boscia, F.; Elkjaer, M.L.; Illes, Z.; Kukley, M. Altered expression of ion channels in white matter lesions of progressive multiple sclerosis: What do we know about their function? Front. Cell. Neurosci. 2021, 15, 685703. [Google Scholar] [CrossRef]
  71. Schultz, C.; Engelhardt, M. Anatomy of the hippocampal formation. Front. Neurol. Neurosci. 2014, 34, 6–17. [Google Scholar]
  72. Ito, H.T.; Schuman, E.M. Functional division of hippocampal area CA1 via modulatory gating of entorhinal cortical inputs. Hippocampus 2012, 22, 372–387. [Google Scholar] [CrossRef] [Green Version]
  73. Sun, X.C.; Li, L.; Zhang, M.; Li, W.B.; Li, Q.J.; Zhao, L. Division of CA1, CA3 and DG regions of the hippocampus of Wistar rat. Zhongguo Ying Yong Sheng Li Xue Za Zhi 2012, 28, 189–192. [Google Scholar]
  74. Watson, C.; Binks, D. Elongation of the CA1 field of the septal hippocampus in ungulates. J. Comp. Neurol. 2019, 527, 818–832. [Google Scholar] [CrossRef]
  75. Van Groen, T.; Wyss, J.M. Extrinsic projections from area CA1 of the rat hippocampus: Olfactory, cortical, subcortical, and bilateral hippocampal formation projections. J. Comp. Neurol. 1990, 302, 515–528. [Google Scholar] [CrossRef]
  76. De La Rosa-Prieto, C.; Ubeda-Banon, I.; Mohedano-Moriano, A.; Pro-Sistiaga, P.; Saiz-Sanchez, D.; Insausti, R.; Martinez-Marcos, A. Subicular and CA1 hippocampal projections to the accessory olfactory bulb. Hippocampus 2009, 19, 124–129. [Google Scholar] [CrossRef]
  77. Bliss, T.V.; Cooke, S.F. Long-term potentiation and long-term depression: A clinical perspective. Clinics 2011, 66 (Suppl. 1), 3–17. [Google Scholar] [CrossRef] [Green Version]
  78. Bliss, T.V.; Lomo, T. Long-lasting potentiation of synaptic transmission in the dentate area of the anaesthetized rabbit following stimulation of the perforant path. J. Physiol. 1973, 232, 331–356. [Google Scholar] [CrossRef]
  79. Kemp, A.; Manahan-Vaughan, D. Hippocampal long-term depression and long-term potentiation encode different aspects of novelty acquisition. Proc. Natl. Acad. Sci. USA 2004, 101, 8192–8197. [Google Scholar] [CrossRef] [Green Version]
  80. Hirano, T. Long-term depression and other synaptic plasticity in the cerebellum. Proc. Jpn. Acad. Ser. B Phys. Biol. Sci. 2013, 89, 183–195. [Google Scholar] [CrossRef] [Green Version]
  81. Wiegert, J.S.; Pulin, M.; Gee, C.E.; Oertner, T.G. The fate of hippocampal synapses depends on the sequence of plasticity-inducing events. Elife 2018, 7, e39151. [Google Scholar] [CrossRef]
  82. Sakurai, M. Synaptic modification of parallel fibre-Purkinje cell transmission in in vitro guinea-pig cerebellar slices. J. Physiol. 1987, 394, 463–480. [Google Scholar] [CrossRef]
  83. Wiegert, J.S.; Oertner, T.G. Long-term depression triggers the selective elimination of weakly integrated synapses. Proc. Natl. Acad. Sci. USA 2013, 110, E4510–E4519. [Google Scholar] [CrossRef] [Green Version]
  84. Lezmy, J.; Gelman, H.; Katsenelson, M.; Styr, B.; Tikochinsky, E.; Lipinsky, M.; Peretz, A.; Slutsky, I.; Attali, B. M-current inhibition in hippocampal excitatory neurons triggers intrinsic and synaptic homeostatic responses at different temporal scales. J. Neurosci. 2020, 40, 3694–3706. [Google Scholar] [CrossRef]
  85. Petrovic, M.M.; Nowacki, J.; Olivo, V.; Tsaneva-Atanasova, K.; Randall, A.D.; Mellor, J.R. Inhibition of post-synaptic Kv7/KCNQ/M channels facilitates long-term potentiation in the hippocampus. PLoS ONE 2012, 7, e30402. [Google Scholar] [CrossRef]
  86. Huang, P.; Li, C.; Fu, T.; Zhao, D.; Yi, Z.; Lu, Q.; Guo, L.; Xu, X. Flupirtine attenuates chronic restraint stress-induced cognitive deficits and hippocampal apoptosis in male mice. Behav. Brain Res. 2015, 288, 1–10. [Google Scholar] [CrossRef]
  87. Stanton, P.K. LTD, LTP, and the sliding threshold for long-term synaptic plasticity. Hippocampus 1996, 6, 35–42. [Google Scholar] [CrossRef]
  88. McCutchen, E.; Scheiderer, C.L.; Dobrunz, L.E.; McMahon, L.L. Coexistence of muscarinic long-term depression with electrically induced long-term potentiation and depression at CA3-CA1 synapses. J. Neurophysiol. 2006, 96, 3114–3121. [Google Scholar] [CrossRef] [Green Version]
  89. Milner, A.J.; Cummings, D.M.; Spencer, J.P.; Murphy, K.P. Bi-directional plasticity and age-dependent long-term depression at mouse CA3-CA1 hippocampal synapses. Neurosci. Lett. 2004, 367, 1–5. [Google Scholar] [CrossRef]
  90. Fontán-Lozano, A.; Suárez-Pereira, I.; Delgado-García, J.M.; Carrión, A.M. The M-current inhibitor XE991 decreases the stimulation threshold for long-term synaptic plasticity in healthy mice and in models of cognitive disease. Hippocampus 2011, 21, 22–32. [Google Scholar] [CrossRef]
  91. Milh, M.; Roubertoux, P.; Biba, N.; Chavany, J.; Spiga Ghata, A.; Fulachier, C.; Collins, S.C.; Wagner, C.; Roux, J.C.; Yalcin, B.; et al. A knock-in mouse model for KCNQ2-related epileptic encephalopathy displays spontaneous generalized seizures and cognitive impairment. Epilepsia 2020, 61, 868–878. [Google Scholar] [CrossRef] [Green Version]
  92. Baculis, B.C.; Zhang, J.; Chung, H.J. The role of Kv7 channels in neural plasticity and behavior. Front. Physiol. 2020, 11, 568667. [Google Scholar] [CrossRef]
  93. Thomann, P.A.; Seidl, U.; Brinkmann, J.; Hirjak, D.; Traeger, T.; Wolf, R.C.; Essig, M.; Schroder, J. Hippocampal morphology and autobiographic memory in mild cognitive impairment and Alzheimer’s disease. Curr. Alzheimer Res. 2012, 9, 507–515. [Google Scholar] [CrossRef]
  94. Hirjak, D.; Wolf, R.C.; Remmele, B.; Seidl, U.; Thomann, A.K.; Kubera, K.M.; Schröder, J.; Maier-Hein, K.H.; Thomann, P.A. Hippocampal formation alterations differently contribute to autobiographic memory deficits in mild cognitive impairment and Alzheimer’s disease. Hippocampus 2017, 27, 702–715. [Google Scholar] [CrossRef]
  95. Li, X.T. Alzheimer’s disease therapy based on acetylcholinesterase inhibitor/blocker effects on voltage-gated potassium channels. Metab. Brain Dis. 2022, 37, 581–587. [Google Scholar] [CrossRef]
  96. Spoleti, E.; Krashia, P.; La Barbera, L.; Nobili, A.; Lupascu, C.A.; Giacalone, E.; Keller, F.; Migliore, M.; Renzi, M.; D’Amelio, M. Early derailment of firing properties in CA1 pyramidal cells of the ventral hippocampus in an Alzheimer’s disease mouse model. Exp. Neurol. 2021, 350, 113969. [Google Scholar] [CrossRef]
  97. Moriguchi, S.; Inagaki, R.; Fukunaga, K. Memantine improves cognitive deficits via KATP channel inhibition in olfactory bulbectomized mice. Mol. Cell. Neurosci. 2021, 117, 103680. [Google Scholar] [CrossRef]
  98. Djebari, S.; Iborra-Lázaro, G.; Temprano-Carazo, S.; Sánchez-Rodríguez, I.; Nava-Mesa, M.O.; Múnera, A.; Gruart, A.; Delgado-García, J.M.; Jiménez-Díaz, L.; Navarro-López, J.D. G-Protein-gated inwardly rectifying potassium (Kir3/GIRK) channels govern synaptic plasticity that supports hippocampal-dependent cognitive functions in male mice. J. Neurosci. 2021, 41, 7086–7102. [Google Scholar] [CrossRef]
  99. Ashrafuzzaman, M. Mitochondrial ion channels in aging and related diseases. Curr. Aging Sci. 2022. [Google Scholar] [CrossRef]
  100. Islas, Á.A.; Scior, T.; Torres-Ramirez, O.; Salinas-Stefanon, E.M.; Lopez-Lopez, G.; Flores-Hernandez, J. Computational molecular characterization of the interaction of acetylcholine and the NMDA receptor to explain the direct glycine-competitive potentiation of NMDA-mediated neuronal currents. ACS Chem. Neurosci. 2022, 13, 229–244. [Google Scholar] [CrossRef]
  101. Cocozza, G.; Garofalo, S.; Capitani, R.; D’Alessandro, G.; Limatola, C.; Taylor, K.I. Microglial potassium channels: From homeostasis to neurodegeneration. Biomolecules 2021, 11, 1774. [Google Scholar] [CrossRef]
  102. Sharma, K.; Pradhan, S.; Duffy, L.K.; Yeasmin, S.; Bhattarai, N.; Schulte, M.K. Role of receptors in relation to plaques and tangles in Alzheimer’s disease pathology. Int. J. Mol. Sci. 2021, 22, 12987. [Google Scholar] [CrossRef]
  103. Hirni, D.I.; Kivisaari, S.L.; Monsch, A.U.; Taylor, K.I. Distinct neuroanatomical bases of episodic and semantic memory performance in Alzheimer’s disease. Neuropsychologia 2013, 51, 930–937. [Google Scholar] [CrossRef] [Green Version]
  104. Bomilcar, I.; Bertrand, E.; Morris, R.G.; Mograbi, D.C. The seven selves of dementia. Front. Psychiatry 2021, 12, 646050. [Google Scholar] [CrossRef]
  105. Liu, H.; Jia, L.; Chen, X.; Shi, L.; Xie, J. The Kv7/KCNQ channel blocker XE991 protects nigral dopaminergic neurons in the 6-hydroxydopamine rat model of Parkinson’s disease. Brain Res. Bull. 2018, 137, 132–139. [Google Scholar] [CrossRef]
  106. Sibille, K.T.; Bartsch, F.; Reddy, D.; Fillingim, R.B.; Keil, A. Increasing neuroplasticity to bolster chronic pain treatment: A role for intermittent fasting and glucose administration? J. Pain. 2016, 17, 275–281. [Google Scholar] [CrossRef] [Green Version]
  107. Paventi, G.; Virginia Soldovieri, M.; Servettini, I.; Barrese, V.; Miceli, F.; Josè Sisalli, M.; Ambrosino, P.; Mosca, I.; Vinciguerra, I.; Testai, L.; et al. Kv7.4 channels regulate potassium permeability in neuronal mitochondria. Biochem. Pharmacol. 2022, 197, 114931. [Google Scholar] [CrossRef]
  108. Bloms-Funke, P.; Bankstahl, M.; Bankstahl, J.; Kneip, C.; Schröder, W.; Löscher, W. The novel dual-mechanism Kv7 potassium channel/TSPO receptor activator GRT-X is more effective than the Kv7 channel opener retigabine in the 6-Hz refractory seizure mouse model. Neuropharmacology 2022, 203, 108884. [Google Scholar] [CrossRef]
  109. Clark, S.; Antell, A.; Kaufman, K. New antiepileptic medication linked to blue discoloration of the skin and eyes. Ther. Adv. Drug Saf. 2015, 6, 15–19. [Google Scholar] [CrossRef] [Green Version]
  110. Wang, J.; Liu, Y.; Hu, F.; Yang, J.; Guo, X.; Hou, X.; Ju, C.; Wang, K. Activation of neuronal voltage-gated potassium Kv7/KCNQ/M-current by a novel channel opener SCR2682 for alleviation of chronic pain. J. Pharmacol. Exp. Ther. 2021, 377, 20–28. [Google Scholar] [CrossRef]
  111. Rivera-Arconada, I.; Vicente-Baz, J.; Lopez-Garcia, J.A. Targeting Kv7 channels in pain pathways. Oncotarget 2017, 8, 12554–12555. [Google Scholar] [CrossRef] [Green Version]
  112. Vicente-Baz, J.; Lopez-Garcia, J.A.; Rivera-Arconada, I. Implication of Kv7 channels in the spinal antinociceptive actions of celecoxib. J. Pharmacol. Exp. Ther. 2019, 370, 472–479. [Google Scholar] [CrossRef] [PubMed]
  113. Vicente-Baz, J.; Rivera-Arconada, I. Spinal actions of the NSAID diclofenac on nociceptive transmission in comparison to the Kv7 channel opener flupirtine. Neuroscience 2020, 440, 186–195. [Google Scholar] [CrossRef] [PubMed]
  114. Tariot, P.; Cumming, J.; Martin, M.; Ballard, C.; Ertan-Lyons, D.; Sultzer, D.; Devanand, D.; Weintraub, D.; McEvoy, B.; Youakim, J.; et al. Trial of pimvanserin in dementia-related psychosis. N. Engl. J. Med. 2021, 385, 3099–3319. [Google Scholar] [CrossRef] [PubMed]
  115. Vicente-Baz, J.; Lopez-Garcia, J.A.; Rivera-Arconada, I. Effects of novel subtype selective M-current activators on spinal reflexes in vitro: Comparison with retigabine. Neuropharmacology 2016, 109, 131–138. [Google Scholar] [CrossRef] [PubMed]
  116. Vicente-Baz, J.; Lopez-Garcia, J.A.; Rivera-Arconada, I. Central sensitization of dorsal root potentials and dorsal root reflexes: An in vitro study in the mouse spinal cord. Eur. J. Pain. 2021, 26, 356–369. [Google Scholar] [CrossRef]
  117. Teng, B.C.; Song, Y.; Zhang, F.; Ma, T.Y.; Qi, J.L.; Zhang, H.L.; Li, G.; Wang, K. Activation of neuronal Kv7/KCNQ/M-channels by the opener QO58-lysine and its anti-nociceptive effects on inflammatory pain in rodents. Acta Pharmacol. Sin. 2016, 37, 1054–1062. [Google Scholar] [CrossRef] [Green Version]
  118. Wang, H.R.; Hu, S.W.; Zhang, S.; Song, Y.; Wang, X.Y.; Wang, L.; Li, Y.Y.; Yu, Y.M.; Liu, H.; Liu, D.; et al. KCNQ channels in the mesolimbic reward circuit regulate nociception in chronic pain in mice. Neurosci. Bull. 2021, 37, 597–610. [Google Scholar] [CrossRef]
  119. Chinta, S.J.; Andersen, J.K. Dopaminergic neurons. Int. J. Biochem. Cell Biol. 2005, 37, 942–946. [Google Scholar] [CrossRef]
  120. William, M.; Singh, S.; Chu, X.-P. Commentary: Large acid-evoked currents, mediated by ASIC1a, accompany differentiation in human dopaminergic neurons. Front. Cell. Neurosci. 2021, 15, 789354. [Google Scholar] [CrossRef]
  121. Hayes, M.T. Parkinson’s disease and parkinsonism. Am. J. Med. 2019, 132, 802–807. [Google Scholar] [CrossRef]
  122. Armstrong, M.J.; Okun, M.S. Diagnosis and treatment of Parkinson disease: A review. JAMA 2020, 323, 548–560. [Google Scholar] [CrossRef] [PubMed]
  123. Chen, X.; Xue, B.; Wang, J.; Liu, H.; Shi, L.; Xie, J. Potassium channels: A potential therapeutic target for Parkinson’s disease. Neurosci. Bull. 2018, 34, 341–348. [Google Scholar] [CrossRef] [PubMed]
  124. Hansen, H.H.; Weikop, P.; Mikkelsen, M.D.; Rode, F.; Mikkelsen, J.D. The pan-Kv7 (KCNQ) channel opener retigabine inhibits striatal excitability by direct action on striatal neurons in vivo. Basic Clin. Pharmacol. Toxicol. 2017, 120, 46–51. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Simola, N.; Morelli, M.; Carta, A.R. The 6-hydroxydopamine model of Parkinson’s disease. Neurotox. Res. 2007, 11, 151–167. [Google Scholar] [CrossRef] [PubMed]
  126. Tieu, K. A guide to neurotoxic animal models of Parkinson’s disease. Cold Spring Harb. Perspect. Med. 2011, 1, a009316. [Google Scholar] [CrossRef] [PubMed]
  127. Pandey, S.; Srivanitchapoom, P. Levodopa-induced dyskinesia: Clinical features, pathophysiology, and medical management. Ann. Indian Acad. Neurol. 2017, 20, 190–198. [Google Scholar]
  128. Tronci, E.; Francardo, V. Animal models of L-DOPA-induced dyskinesia: The 6-OHDA-lesioned rat and mouse. J. Neural. Transm. 2018, 125, 1137–1144. [Google Scholar] [CrossRef]
  129. Sander, S.E.; Lemm, C.; Lange, N.; Hamann, M.; Richter, A. Retigabine, a K(V)7 (KCNQ) potassium channel opener, attenuates L-DOPA-induced dyskinesias in 6-OHDA-lesioned rats. Neuropharmacology 2012, 62, 1052–1061. [Google Scholar] [CrossRef]
  130. Sander, S.E.; Lambrecht, C.; Richter, A. The K(V)7.2/3 preferring channel opener ICA 27243 attenuates L-DOPA-induced dyskinesia in hemiparkinsonian rats. Neurosci. Lett. 2013, 545, 59–63. [Google Scholar] [CrossRef]
  131. Hansen, H.H.; Waroux, O.; Seutin, V.; Jentsch, T.J.; Aznar, S.; Mikkelsen, J.D. Kv7 channels: Interaction with dopaminergic and serotonergic neurotransmission in the CNS. J. Physiol. 2008, 586, 1823–1832. [Google Scholar] [CrossRef]
  132. McCutcheon, R.A.; Abi-Dargham, A.; Howes, O.D. Schizophrenia, dopamine and the striatum: From biology to symptoms. Trends Neurosci. 2019, 42, 205–220. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Mattmann, M.E.; Yu, H.; Lin, Z.; Xu, K.; Huang, X.; Long, S.; Wu, M.; McManus, O.B.; Engers, D.W.; Le, U.M.; et al. Identification of (R)-N-(4-(4-methoxyphenyl) thiazol-2-yl)-1-tosylpiperidine-2-carboxamide, ML277, as a novel, potent and selective K(v)7.1 (KCNQ1) potassium channel activator. Bioorg. Med. Chem. Lett. 2012, 22, 5936–5941. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Yu, H.; Lin, Z.; Mattmann, M.E.; Zou, B.; Terrenoire, C.; Zhang, H.; Wu, M.; McManus, O.B.; Kass, R.S.; Lindsley, C.W.; et al. Dynamic subunit stoichiometry confers a progressive continuum of pharmacological sensitivity by KCNQ potassium channels. Proc. Natl. Acad. Sci. USA 2013, 110, 8732–8737. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Leitner, M.G.; Halaszovich, C.R.; Oliver, D. Aminoglycosides inhibit KCNQ4 channels in cochlear outer hair cells via depletion of phosphatidylinositol (4,5) bisphosphate. Mol. Pharmacol. 2011, 79, 51–60. [Google Scholar] [CrossRef] [Green Version]
  136. Abbott, G.W.; Redford, K.E.; Yoshimura, R.F.; Manville, R.W.; Moreira, L.; Tran, K.; Arena, G.; Kookootsedes, A.; Lasky, E.; Gunnison, E. KCNQ and KCNE isoform-dependent pharmacology rationalizes native American dual use of specific plants as both analgesics and gastrointestinal therapeutics. Front. Physiol. 2021, 12, 777057. [Google Scholar] [CrossRef]
  137. Volberg, W.A.; Koci, B.J.; Su, W.; Lin, J.; Zhou, J. Blockade of human cardiac potassium channel human ether-a-go-go-related gene (HERG) by macrolide antibiotics. J. Pharmacol. Exp. Ther. 2002, 302, 320–327. [Google Scholar] [CrossRef] [Green Version]
  138. Yang, T.; McBride, B.F.; Leake, B.F.; Kim, R.B.; Roden, D.M. Modulation of drug block of the cardiac potassium channel KCNA5 by the drug transporters OCTN1 and MDR1. Br. J. Pharmacol. 2010, 161, 1023–1033. [Google Scholar] [CrossRef] [Green Version]
  139. Grünert, S.C.; Bodi, I.; Odening, K.E. Possible mechanisms for sensorineural hearing loss and deafness in patients with propionic acidemia. Orphanet J. Rare Dis. 2017, 12, 30. [Google Scholar] [CrossRef] [Green Version]
  140. Naffaa, M.M.; Al-Ewaidat, O.A. Ligand modulation of KCNQ-encoded (KV7) potassium channels in the heart and nervous system. Eur. J. Pharmacol. 2021, 906, 174278. [Google Scholar] [CrossRef]
  141. Barrese, V.; Stott, J.B.; Baldwin, S.N.; Mondejar-Parreño, G.; Greenwood, I.A. SMIT (Sodium-Myo-Inositol Transporter) 1 regulates arterial contractility through the modulation of vascular Kv7 channels. Arterioscler. Thromb. Vasc. Biol. 2020, 40, 2468–2480. [Google Scholar]
  142. Larsson, J.E.; Frampton, D.J.A.; Liin, S.I. Polyunsaturated fatty acids as modulators of KV7 channels. Front. Physiol. 2020, 11, 641. [Google Scholar] [CrossRef] [PubMed]
  143. Peixoto-Neves, D.; Kanthakumar, P.; Afolabi, J.M.; Soni, H.; Buddington, R.K.; Adebiyi, A. KV7.1 channel blockade inhibits neonatal renal autoregulation triggered by a step decrease in arterial pressure. Am. J. Physiol. Renal Physiol. 2022, 322, F197–F207. [Google Scholar] [CrossRef] [PubMed]
  144. Liu, X.W.; Wu, H.M.; Bai, Y.; Zeng, Q.; Cao, Z.M.; Wu, X.S.; Tang, M. Potassium channel Shaker play a protective role against cardiac aging in Drosophila. Yi Chuan 2021, 43, 94–99. [Google Scholar] [PubMed]
  145. Stott, J.B.; Barrese, V.; Suresh, M.; Masoodi, S.; Greenwood, I.A. Investigating the role of G protein βγ in Kv7-dependent relaxations of the rat vasculature. Arterioscler. Thromb. Vasc. Biol. 2018, 38, 2091–2102. [Google Scholar] [CrossRef] [Green Version]
  146. Blandin, C.E.; Gravez, B.J.; Hatem, S.N.; Balse, E. Remodeling of ion channel trafficking and cardiac arrhythmias. Orphanet J. Rare Dis. 2021, 10, 2417. [Google Scholar] [CrossRef]
  147. Gallego, M.; Zayas-Arrabal, J.; Alquiza, A.; Apellaniz, B.; Casis, O. Electrical features of the diabetic myocardium. Arrhythmic and cardiovascular safety considerations in diabetes. Front. Pharmacol. 2021, 12, 687256. [Google Scholar] [CrossRef]
  148. Herrera-Pérez, S.; Campos-Ríos, A.; Rueda-Ruzafa, L.; Lamas, J.A. Contribution of K2P potassium channels to cardiac physiology and pathophysiology. Int. J. Mol. Sci. 2021, 22, 6635. [Google Scholar] [CrossRef]
  149. Yang, K.; Xue, Y.; Yu, M.; Jiao, H.; Li, Y.; Wei, X.; Liu, W.; Sun, Y.; Chen, N.; Song, L.; et al. Protective effect of trimetazidine on potassium ion homeostasis in myocardial tissue in mice with heart failure. Biomed. Res. Int. 2022, 2022, 2387860. [Google Scholar] [CrossRef]
Figure 1. KCNQ channel structure is composed of six transmembrane segmental helices, classified as S1–S6. In addition, the S1–S4 transmembrane segmental helices compose the actual voltage sensation region, and the latter two (S5–S6) units are the actual gate of the channel.
Figure 1. KCNQ channel structure is composed of six transmembrane segmental helices, classified as S1–S6. In addition, the S1–S4 transmembrane segmental helices compose the actual voltage sensation region, and the latter two (S5–S6) units are the actual gate of the channel.
Membranes 12 00499 g001
Figure 2. KCNQ channels and pharmacotherapeutic modulators in relation to the cellular membrane. Retigabine acts to open the KCNQ channel, whereas linopirdine and XE991 are channel inhibitors that act to inactivate the KCNQ channel function.
Figure 2. KCNQ channels and pharmacotherapeutic modulators in relation to the cellular membrane. Retigabine acts to open the KCNQ channel, whereas linopirdine and XE991 are channel inhibitors that act to inactivate the KCNQ channel function.
Membranes 12 00499 g002
Table 1. Expression distribution and associated pathologies with channel genes.
Table 1. Expression distribution and associated pathologies with channel genes.
GeneExpression DistributionAssociated Pathologies
KCNQ1CochleaType 1 long QT syndrome
Heart
KCNQ2CerebellumBenign familial neonatal seizures
Hippocampus
Medulla
Early onset epileptic encephalopathy
KCNQ3CerebellumBenign familial neonatal seizures
Hippocampus
Medulla
Early onset epileptic encephalopathy
Bipolar Disorder
KCNQ4Cochlea
Trigeminal ganglia
Deafness
KCNQ5Retinal pigment epithelium*
* No major associated pathologies. Of note, this table is not comprehensive to all expression and pathological distributions of these genes.
Table 2. Current Proposed Neuromodulation Tactics and Agents utilizing KCNQ Channels.
Table 2. Current Proposed Neuromodulation Tactics and Agents utilizing KCNQ Channels.
Action on ChannelMechanismsAgents
OpeningPhosphatidyl inositol 4, 5 biphosphate (PIP 2)
Gamma-amino butyric acid (GABA)
Inositol 1, 4, 5 triphosphate (IP3)
a. PIP2 synthesis
b. Suppression of muscarinic current (M) receptor at acetyl choline receptor
Retigabine
Flupirtine
ICA 27243
ClosingMuscarinic current (M) associated with acetyl choline receptor
Pathological Mechanisms
a. Movement Disorders
Linoperidine
XE991
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Singh, S.P.; William, M.; Malavia, M.; Chu, X.-P. Behavior of KCNQ Channels in Neural Plasticity and Motor Disorders. Membranes 2022, 12, 499. https://doi.org/10.3390/membranes12050499

AMA Style

Singh SP, William M, Malavia M, Chu X-P. Behavior of KCNQ Channels in Neural Plasticity and Motor Disorders. Membranes. 2022; 12(5):499. https://doi.org/10.3390/membranes12050499

Chicago/Turabian Style

Singh, Som P., Matthew William, Mira Malavia, and Xiang-Ping Chu. 2022. "Behavior of KCNQ Channels in Neural Plasticity and Motor Disorders" Membranes 12, no. 5: 499. https://doi.org/10.3390/membranes12050499

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