Next Article in Journal
CBL/CAP Is Essential for Mitochondria Respiration Complex I Assembly and Bioenergetics Efficiency in Muscle Cells
Next Article in Special Issue
Aromatic Bromination Abolishes Deficits in Visuospatial Learning Induced by MDMA (“Ecstasy”) in Rats While Preserving the Ability to Increase LTP in the Prefrontal Cortex
Previous Article in Journal
Mechanical, Barrier and Thermal Properties of Amylose-Argan Proteins-Based Bioplastics in the Presence of Transglutaminase
Previous Article in Special Issue
Obesogenic Diet-Induced Neuroinflammation: A Pathological Link between Hedonic and Homeostatic Control of Food Intake
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Ca2+- and Voltage-Activated K+ (BK) Channels in the Nervous System: One Gene, a Myriad of Physiological Functions

by
Carlos Ancatén-González
1,2,
Ignacio Segura
1,
Rosangelina Alvarado-Sánchez
1,3,
Andrés E. Chávez
1,*,† and
Ramon Latorre
1,*,†
1
Centro Interdisciplinario de Neurociencia de Valparaíso (CINV), Instituto de Neurociencias, Universidad de Valparaíso, Valparaíso 2340000, Chile
2
Programa de Doctorado en Ciencias, Mención Neurociencia, Universidad de Valparaíso, Valparaíso 2340000, Chile
3
Doctorado en Ciencias Mención Biofísica y Biología Computacional, Universidad de Valparaíso, Valparaíso 2340000, Chile
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2023, 24(4), 3407; https://doi.org/10.3390/ijms24043407
Submission received: 19 December 2022 / Revised: 18 January 2023 / Accepted: 21 January 2023 / Published: 8 February 2023
(This article belongs to the Special Issue State-of-the-Art Molecular Neurobiology in Chile)

Abstract

:
BK channels are large conductance potassium channels characterized by four pore-forming α subunits, often co-assembled with auxiliary β and γ subunits to regulate Ca2+ sensitivity, voltage dependence and gating properties. BK channels are abundantly expressed throughout the brain and in different compartments within a single neuron, including axons, synaptic terminals, dendritic arbors, and spines. Their activation produces a massive efflux of K+ ions that hyperpolarizes the cellular membrane. Together with their ability to detect changes in intracellular Ca2+ concentration, BK channels control neuronal excitability and synaptic communication through diverse mechanisms. Moreover, increasing evidence indicates that dysfunction of BK channel-mediated effects on neuronal excitability and synaptic function has been implicated in several neurological disorders, including epilepsy, fragile X syndrome, mental retardation, and autism, as well as in motor and cognitive behavior. Here, we discuss current evidence highlighting the physiological importance of this ubiquitous channel in regulating brain function and its role in the pathophysiology of different neurological disorders.

1. Introduction

Large-conductance calcium (Ca2+)- and voltage-activated potassium (K+) channels (BK, also known as Maxi-K or Slo1) are ubiquitously expressed in the body playing important roles in blood flow regulation, renal excretion, muscle contraction, circadian rhythm, and hearing [1,2,3]. BK channels are also abundantly expressed throughout the central nervous system in the cortex, basal ganglia, hippocampus, thalamus, and cerebellum, among other areas (Table 1) [4,5,6]. There, BK channels are strategically positioned on the plasma membrane and intracellular compartments in both neurons and glial cells [7,8,9,10,11,12,13]. Typically, activation of BK channels leads to a massive efflux of K+ ions that hyperpolarizes cellular membrane potential to regulate neuronal excitability and synaptic function, and, ultimately, motor and cognitive behavior [14,15,16,17]. Within a single neuron, BK channels are known to be expressed within different compartments, accomplishing different functions [18], including somatic regulation of action potentials [10,19] and regulation of neurotransmitter release at synaptic terminals [16,20,21,22,23,24]. BK channels are also present in dendritic compartments and neuronal spines [25], where they can limit the postsynaptic potential [25,26,27] or dendritic Ca2+ signals [25,28]. Although BK channels are also localized in nuclei, lysosomes, and mitochondria [11], their contribution to regulating neuronal function remains elusive.
While increasing evidence supports a critical role for BK channels in regulating neuronal excitability, their ability to act as a coincidence detector for changes in intracellular Ca2+ concentration suggests that BK channels are also well-positioned to control synaptic communication by exerting either negative [23,24] or positive [29] regulation of Ca2+ influx through voltage-activated Ca2+ channels (for review see Griguoli et al., 2016 [16]). However, the mechanisms by which BK regulate neuronal function and behavior in the brain remain far from being completely understood.

2. BK Biophysical and Structural Characteristics

Coded for by a single gene (Slo1, KCNMA1), BK channels have been proven, from the very beginning, to show a complex, albeit very attractive behavior. Their large conductance (≈250 pS in symmetrical 100 mM K+) near the limit of the diffusional process of K+ in solution is accompanied by an exquisite K+ selectivity, making them unique in the large family of different K+ channels [30,31,32]. Cloning of BK channels showed their similarity to voltage-dependent K+ channels in which each subunit contains six transmembrane domains, including a well-defined voltage sensor (S1–S4) and a pore domain (S5–S6) [33,34]. However, later studies showed that the channel-forming protein contained an extra segment (S0) placing the N-terminus on the extracellular side [35]. In addition, BK possess a large C-terminus, composed of two regulators of K+ conductance (RCK1 and RCK2), where the high-affinity Ca2+-binding sites reside [36,37,38,39].
Once single-channel, macroscopic, and gating currents records were available, one of the main challenges was to combine BK voltage dependence and channel activation by Ca2+ in one gating-kinetics scheme. From the start, it became clear that BK gating kinetics are intricate. One big step ahead was achieved by the group of Magleby [40], who identified that the several open and closed states present in the single-channel current records were correlated and inversely related. Shorter open intervals were followed by long closed intervals and vice versa. These observations made linear gating-kinetics models as those proposed, for example, for shaker K+ channels, unlikely. Instead, they suggested a progression of closed states of decreasing duration connected to open states of increasing lifetimes. More importantly, these findings hinted that the gating in BK channels is allosteric. Another important piece of information was that treatment of BK channels with N-bromoacetamide rendered BK voltage-dependent channels insensitive to internal Ca2+, a clear indication that voltage sensors and Ca2+-binding sites were in different regions of the channel-forming protein [41]. These early discoveries allowed us to define BK channels from the start as modular proteins with allosteric gating before knowing their well-defined structure.
The BK channel gating characteristics have been summarized in a two-tiered kinetic model [42,43]. Considering the tetrameric nature of BK channels in this allosteric gating-kinetics model, the four voltage sensor domains (VSD) undergo a transition between resting (R) and active states (A). The transition R–A is defined by a voltage-dependent equilibrium constant J. The VSDs allosterically interact with the pore opening through the allosteric factor D (Figure 1A). The model considers the existence of four Ca2+-binding sites, each of which includes a Ca2+-binding mediated transition between an empty site (U) and a Ca2+-bound site (BCa2+) with an equilibrium constant K = [Ca2+]/KD. In this case, the allosteric interaction with the pore opening is determined by the allosteric factor C. On the other hand, the strength of allosteric interaction between voltage sensors and Ca2+ binding is determined by the allosteric factor E. It is important to note here that the closed–open (C–O) reaction defined by the equilibrium constant L can proceed in the absence of voltage-sensor activation and of internal Ca2+, albeit with a very low probability of opening (Figure 1B). Figure 2B shows that Ca2+ alone can increase the open probability (Po) by about four orders of magnitude. Since Po increases with depolarizing voltages and internal Ca2+ concentrations, and BK has large conductance, BK channels are the perfect damping machine to brake excitatory processes mediated by internal Ca2+.
There is a strong coupling between the Ca2+ binding sites and the voltage sensors. At saturating Ca2+ concentrations, the equilibrium constant J that defines the resting–active equilibrium of the voltage sensor (Figure 1A) increases 26-fold [44,45]. This result implies that the free energy necessary to activate the BK channel voltage sensor decreases by about 8 kJ/mol at saturating internal Ca2+ concentrations (100 µM).
The modular nature of BK channels was confirmed when the first BK structures were determined using cryo-microscopy [38,46]. As voltage-dependent K+ (Kv) channels, BK channels are tetramers composed of four identical α subunits (Figure 2A), but unlike Kv channels, BK channels have a non-swapped configuration. The VSD of one subunit (S1–S4) contacts the pore domain (S5–Pore domain–S6) of the same subunit (Figure 2 A, C). Although the work of Ma et al. [47] suggested the presence of a decentralized VSD in which the gating charges were spread across S2, S3, and S4, more recent work gave strong evidence that the gating charges are two arginines contained in S4 (R210 and R213) located in a septum devoid of water [48]. The movements of the charges during BK-VSD activation are small, with only a modest (≈2 Å) displacement of S4. It is unclear at present how the electrical energy is transformed into opening of the pore. However, the gating ring probably mediates the coupling between VSDs and the pore [38,45,46].
The two RCK domains form a gating ring (Figure 1B) containing the two high-affinity Ca2+-binding domains that can be distinguished structurally by their selectivity to divalent ions. The divalent cation-binding site in RCK1 preferentially binds Ca2+ and Cd2+, whereas in RCK2, the preferred divalent cations are Ca2+ and Ba2+ [38,49,50]. The binding of Ca2+ promotes an expansion of the gating ring, which, by pulling the linker that connects the gating ring to S6, causes this transmembrane segment to adopt a configuration that leads to channel opening [46]. It is unclear at present where the gate that defines the closed–open transition is located (Figure 1A). The BK structures show that there is only a modest decrease in the radius of the internal BK vestibule when the channel closes, which varies between 1 and 4 Å [38]. The diameter of the narrowest part of the vestibule is 12 Å. If there is no well-defined structure hindering ions’ passage through the BK channel, the following question arises: Where is the gate? Blockade mediated by large quaternary ammonium ions gave the first hint suggesting that, similarly to other channels [51,52], the BK gate resides in a selectivity filter [53,54,55]. However, simulations of molecular dynamics using BK structures in the absence and the presence of Ca2+ indicated that the actual decrease in diameter of the internal cavity of the BK channel during closed state is larger than the one shown by the BK structure obtained in the absence of Ca2+ [56]. The final diameter of the cavity was 6–8 Å, and its walls became more hydrophobic, creating a region devoid of water. Due to its low dielectric constant, it produces an energy barrier unsurmountable to ions.

3. BK Auxiliary Subunits

The four auxiliary β subunits (β1–β4) in BK are characterized by a common structure consisting of two transmembrane segments connected by a large extracellular loop (~148 aa residues) and cytoplasmic COOH and NH2 termini. Although to a different degree, all β subunits modify BK channel gating and pharmacology (reviewed in [10,57,58]), β1, β2, and β4 increase Ca2+ sensitivity by stabilizing the voltage sensor’s active conformation and slowing down the activation and deactivation kinetics. Mediated by its NH2 terminus, β2 also induces a fast and complete inactivation. Four isoforms are generated by splicing the β3 gene (kcnmb3; β3a–d). β3 subunits do not modify BK channel Ca2+ sensitivity, but three of them (β3a–c) produce a fast and incomplete inactivation. Importantly, β subunits have different pharmacological profiles, as β2 is iberiotoxin (IBTX)-sensitive, whereas β4 is IBTX-resistant. However, BK channels consisting of the complexes α/β2 or α/β4 can be blocked by paxilline [10,57,59,60]. β4 slows down the activation and deactivation kinetics at all Ca2+ concentrations. In symmetrical K+-concentration conditions, β4 promotes a leftward shift of the BK conductance–voltage curve only at Ca2+ concentrations >10 µM compared to BK channels formed by the α subunit alone [61,62]. However, here it is essential to highlight the fact that under physiological conditions—low external K+—BK/β4 channels shift the conductance–voltage curves leftward at all internal Ca2+ concentrations compared to BK/α channels [63,64].
When characterizing the different types of Ca2+-activated channels in the brain plasma membrane vesicles, Reinhart et al. [65] detected two types of high-conductance BK channels. One type showed fast gating and was blocked by charybdotoxin (ChTX) (type I), and the other was ChTX-resistant and exhibited slow gating (type II). Cloning of the different β subunits (β1–β4) showed that the toxin resistance of some BK channels in the brain was most probably due to the formation of complexes between the pore-forming α and the β4 subunit [60,66,67]. Swapping the long external loop connecting TM1 and TM2 between β4 and β1 and neutralization of some of its basic residues showed that this region of the β4 subunit hinders the binding of iberiotoxin (IBTX) by blocking access of the toxin to the BK external vestibule [60,68]. The cryo-structures of the α/β4 BK channel show a 1:1 stoichiometry between α and β4 subunits in which the four external loops of the β4 give origin to a crown-like structure [69]. The crown impedes both access to and toxin release from the binding site, explaining the extremely slow ON and OFF rates of IBTX-inhibition kinetics [70].
β4 is the most abundant auxiliary subunit expressed in the nervous system [66,71]. β4 is found in the thalamus, brainstem, posterior pituitary terminals, pyramidal neurons of the cortex, CA3 pyramidal neurons in the hippocampus, hippocampal dentate granule cells, olfactory bulb, and Purkinje cerebellar neurons [64,72,73,74,75] (Table 1). β2 is also extensively expressed in the brain, albeit at lower levels than β4 [66,71]. β1 on the other hand, is almost absent in the brain, but it has been detected in the hypothalamus, paraventricular neurons, and cerebellar Purkinje cells [76,77,78]. Interestingly, these hypothalamic neurons reportedly express β1 subunits on the soma and dendrites, while the axon terminals mainly contain the β4 subunit [79]. Such specific distribution of β subunits within the same neurons may have important functional significance in understanding how BK channels differentially contribute to the regulation of somatic and nerve terminal compartments.
The physiological role of the α/β4 BK channels in the brain was first elucidated by Brenner et al. [72], who showed that this type of BK channel (type II) is present in the dentate gyrus granular cells regulating their excitability. Knockout of the β4 converted type II into type I (fast) BK channels [72]. Type I channels reduce, and type II BK channels increase the duration of action potentials [63,72]. Interestingly, in pathological conditions, such as a neuropathic pain model, β4 seems to be upregulated in anterior cingulate cortex pyramidal neurons [73], suggesting that the expression of BK channel subunits could be modulated in an activity-dependent and cellular-specific fashion. Moreover, evidence suggests that the fragile mental retardation protein (FMRP), by interacting with the β4 subunit, shortens the action potential, thus regulating neurotransmitter release in CA3 pyramidal neurons [15]. Single-channel recording in CA3 pyramidal neurons showed that the α/β4 BK channel open probability is reduced by about 50% in neurons from fmpr knockout mice increasing the inter-burst interval [80]. However, a recent study showed that FMPR modulates both α/β4 BK and α BK channels [81]. The degree of importance in modulating synaptic activity via interaction of FMPR with α/β4 BK or α BK channels is currently unclear.
Inactivating BK currents have been detected in mammalian auditory hair cells, rat lateral amygdala neurons, and hippocampal CA1 neurons [82,83,84]. In CA1 and dorsal root ganglion neurons, inactivation is removed when the inner membrane surface is treated with trypsin, suggesting that BK channels are formed by the α/β2 complex [84,85]. Consistent with this idea, in rat dorsal root ganglion neurons, α/β2 BK channels are blocked by IBTX, resulting in increasing firing frequency, broadening of the action potential, and reducing afterhyperpolarization [85]. The spontaneous firing rate in the suprachiasmatic nucleus (SCN) exhibits daily oscillations that determine the normal circadian timing of behavior in mammals. The intrinsic circadian clock in the SCN controls the daily expression of BK. Consistently, BK currents are lower during the day and high at night [86,87,88], thus determining the daily SCN rate of firing. Moreover, inactivating BK currents are predominant during the day and are reduced at night with an increase in steady-state BK currents [86]. Importantly, in the SCN, BK inactivation is mediated by the β2 subunit, and β2 knockout mice show a lack of diurnal variation in BK currents and SCN firing rate [86].

4. BK Channel Coupling with Ca2+-Permeable Channels

CaV and BK Nanodomains: Critical for the feedback control of Ca2+ influx and cell excitability in the nervous system, BK channels co-localize with voltage-dependent Ca2+ (CaV) channels and N-methyl-D-aspartate receptors (NMDAR; reviewed in [89,90]). Due to the low affinity of BK for Ca2+ (Kd ≥ 3 µM; [91]), co-localization with Ca2+-permeable channels provides BK channels with an effective local Ca2+ concentration for their activation and function [92,93]. Indeed, BK channels have been reported to form nanodomains with L-type, P/Q, N-type, R-type, or T-type CaV [94,95,96,97,98].
Ca2+ chelator EGTA and fast chelator BAPTA differ in their Ca2+-binding rates but have similar Ca2+ affinities. They have been a powerful tool for determining the diffusion distance for Ca2+ from calcium channels to BK channels. BK channel activity was abolished when using BAPTA as the internal Ca2+ buffer, but not with EGTA, revealing the close spatial proximity (~10 to 20 nm) in the nanodomains between BK channels and voltage-sensitive calcium (CaV) channels [96,99,100]. More recently, a mathematical model revealed the coexistence of L-type and BK-type CaV channels and suggested that they are grouped in nanodomains spatially separated 30 nm from each other [101].
Activation of BK channels mediated by the influx of Ca2+ through CaV channels decreases the action potential duration, promotes fast hyperpolarization potentials, decreases the time of Ca2+ influx, and limits the release of neurotransmitters [99,102,103]. The activation of CaV channels modifies the voltage dependency, current amplitude, and kinetics of BK channel activation, causing functional variability in BK dependent on the expression pattern of CaV channels [96,104]. The pharmacological blockade of BK channels using IBTX, ChTX, or paxilline has also demonstrated the functional coupling between BK and CaV channels. The application of these toxins increases the amplitude of the action potential and Ca2+ entry through CaV channels as well as the release of neurotransmitters [99,102,105,106,107,108]. On the other hand, the specific blockade of a subset of CaV channels suppresses the hyperpolarization that is provided by the flow of K+ ions from active BK channels [96,97,102,105,106,109,110]. One example that illustrates the importance of CaV and BK localization is the functional and spatial localization of CaV1.3 and BK channels in rat hippocampal and sympathetic neurons [104]. CaV1.3 activates at low voltages, and when co-expressed with BK, the latter started activating at about −50 mV. In hippocampal and sympathetic neurons, CaV1.3 and BK are localized as clusters of CaV1.3 encircling clusters of BK channels, forming a multi-channel complex. Since BK channels activate at voltages near the action potential threshold in these complexes, they enable the modulation of neuronal excitability.
BK and NMDAR Complexes: BK channels and NMDARs are widely expressed in the brain, and subcellular distribution variations are caused by the appearance of distinct BK channel auxiliary subunits and NMDAR subunit composition. So far, the BK–GluN1 complex has been functionally described on granule cells of the olfactory bulb [111], on mature dentate granule cells in the hippocampus [27], and in a subset of layer 5 pyramidal neurons of the barrel cortex and visual cortex [25]. Moreover, pulling-down assays demonstrated that this complex is also present in the cerebellum, cortex, thalamus, and striatum [27].
Functional coupling between NMDARs and BK channels was first demonstrated in outside-out patches from the cell bodies of granule cells, where the glutamate-induced outward currents were abolished after applying TEA, paxilline, or IBTX [111]. Distinctively, in the BK–GluN1 complex, the BK S0–S1 loop interacts with the C-terminal domain of the NMDARs. NMDAR activation by glutamate produced robust BK outward currents in whole-cell voltage-clamp recordings on mature dentate gyrus granule cells [27]. The application of NMDAR antagonist (2R)-amino-5-phosphonovaleric acid (AP5) and the pore-blocker MK-801 inhibited the outward currents, confirming the tight association of the BK–GluN1 complex. BK channel outward currents were also elicited by NMDAR activation in neocortex pyramidal neurons [112], those BK currents were also eliminated by AP5, confirming that NMDAR activation is needed to trigger the opening of BK channels. Likewise, ZnCl2 and ifenprodil (IFEN) were used to inhibit GluN2A or GluN2B-containing NMDARs, respectively, to assess the selective coupling between BK channels and different NMDAR subunit compositions. Nevertheless, these compounds only partially suppressed the BK outward currents, suggesting that GluN2A or GluN2B might also be a part of the functional coupling between BK channels and NMDARs in the brain.
NMDAR–BK coupling efficiency with the different NMDAR subunits was also tested in HEK cells expressing BK and GluN1/GluN2A or GluN1/GluN2B using a proximity ligation assay. In agreement with the pharmacological data, BK channels showed no preference for different NMDAR subunits [112]. To further evaluate the degree of coupling between BK channel activation and NMDAR, the Ca2+ chelators EGTA and BAPTA were used to determine the relative distance between BK channels and NMDARs in granule cell bodies [111]. The BK channel outward currents induced by glutamate were insensitive to the applied intracellular EGTA. In contrast, BAPTA abolished the BK currents induced by the activation of NMDARs. BK–NMDAR coupling has also been tested in hippocampal and pyramidal neurons using Ca2+ chelators [27,112]. In both cases, the results show that BK is robustly activated by the Ca2+ influx mediated by NMDAR, suggesting a short coupling distance.
Other Sources of Ca2+ for BK Channel Activation: Co-localization of BK channels with CaV channels may form a dynamic and reversible channel complex between the plasma membrane and the endoplasmic reticulum. In neurons of the cerebellum, hippocampus, and suprachiasmatic nucleus, the Ca2+ influx mediated by CaV channels activates the ryanodine receptors (RyR) present on a nanodomain of the endoplasmic reticulum. RyR release sparks caused by Ca2+ immediately below the plasmatic membrane, result in generating rapid activation of the BK channels, forming a complex commonly named a CaV–RyR–BK triad [113,114,115]. This triad allows control over action potential firing patterns on a millisecond time scale [113,114]. Moreover, interaction of BK channels with transient receptor potential vanilloid receptor 1 (TRPV1) channels has also been demonstrated in rat dorsal root ganglion neurons and in heterologous systems where both channels were co-expressed. Using electrophysiological recordings and co-immunoprecipitation assays showed that BK channels are activated by Ca2+ influx through TRPV1 channels, with which they associate in complexes in a largely EGTA-insensitive manner [116].

5. BK Post-Translational Modifications

In addition to the diversity of subunits controlling BK function, extensive post-translational modifications significantly modify channel function [117,118,119,120]. These post-translational modifications impact the contribution of BK channels to the control of neuronal function. One of the most studied modifications is the addition of phosphate groups to functionally essential residues (Ser/Thr/Tyr) present within the BK α subunit [9], where at least 30 Ser/Thr phosphorylation sites have been identified. Changes in phosphorylation status differentially modulate the voltage and Ca2+-dependent activation [118]. This effect seems to be developmentally regulated [121] as phosphorylation by protein kinase A (PKA) and protein kinase C (PKC) are inversely predominant during fetal and adult stages [117]. While phosphorylation can suppress BK currents induced by an action potential [122], inhibition of phosphatases or mutation in Ser/Thr phosphorylation sites hinders the slow-down of the BK gating kinetics induced by the β4 subunit [123]. These post-translational modifications of BK channels imply that different functions might arise depending on their phosphorylation status. For instance, in medial vestibular nucleus neurons, BK channels contribute to afterhyperpolarization in a CaMKII-dependent manner [124]. In contrast, PKC decreases BK channel activity [125]. In cerebellar Purkinje neurons, PKA can activate non-inactivating BK channels with low activity but inhibits channels with higher activity [126]. PKCγ on the other hand, negatively modulates BK currents resulting in attenuation of the electrical signal and significant alterations of the complex spike waveform [127]. In presynaptic terminals, activation of BK channels and their subsequent phosphorylation by CaMKII modulates neurotransmitter release [119,128], synaptic depression underlying habituation [129], and input-specific spike-timing-dependent long-term depression at thalamostriatal, but not at corticostriatal synapses [28]. Similar to kinases, phosphatases can also regulate the function of BK channels mainly through an inhibitory control of the channel [106,130,131,132]. Accordingly, type I BK channels lacking β4 can increase neuronal excitability in dentate granule cells in response to reduced phosphatase activity [133].

6. BK Channels Differentially Contribute to Control of Brain Function

The BK α subunit is highly expressed during late embryonic and early postnatal development [14]. As development progresses, BK currents undergo an abrupt increase during the first weeks after birth in neocortical pyramidal neurons, substantia nigra dopamine neurons, and cochlear inner hair cells [134,135,136]. Similarly, BK channel expression in hippocampal neurons is likely to be developmentally regulated, differentially contributing to the repolarization of single action potentials during the first postnatal week [137], suggesting that BK channels contribute to shaping the properties of mature neuronal firing. One of the clearest roles of BK channels in regulating neural activity comes from electrophysiological studies of neuronal firing. Several reports showed that BK channels control action potential shape by contributing to repolarization and fast afterhyperpolarization (fAHP) currents in multiple neuronal types. These include cerebellar Purkinje cells [102,138,139,140], hippocampal pyramidal cells [141,142,143], cortical pyramidal neurons [144,145], and striatal medium spiny neurons [146,147] (Table 1). As discussed above, the strategic co-localization of BK with Cav channels allows BK channels to control Ca2+ influx during action potential repolarization and fAHP, even in spontaneously firing neurons [34], suggesting that BK channels act as a feedback regulator to control neuronal excitability. While the BK–Cav channel coupling indicates that BK channels participate during the repolarization phase of the action potential, their coupling with the RyR contributes to the free Ca2+ necessary for BK channel activation during the fAHP [148]. However, recent evidence in CA1 pyramidal neurons indicates that the BK channel–Cav2.3 complex regulates somatic excitability by controlling fAHP rather than repolarization. Cav2.3 KO mice showed increased excitability with a decreased fAHP [103]. Such differences may arise from the non-homogenous distribution of BK channels in clustered or scattered pools. While clustered pools may be involved in slower events triggered by Ca2+ released by intracellular domains, scattered pools could be involved in transient events such as spike repolarization [139]. Adding to the complexity of BK channel function regulating neuronal firing, evidence indicates that BK channels could also increase neuronal firing [142]. By accelerating membrane repolarization and limiting sodium channel inactivation, BK channels can promote repetitive spiking [149]. An enhanced fAHP can also increase neuronal firing by promoting sodium-channel recovery from inactivation [150].
In some neurons, such as Purkinje [177] and vestibular neurons [178], inhibiting BK channels speeds up action potential firing accompanied by a reduction in afterhyperpolarization. Paradoxically, in substantia nigra dopaminergic neurons, inhibiting BK channels increased spike width. This effect was accompanied by an increase in afterhyperpolarization current, likely involving an enhancement of the slowly deactivating Kv2 current [179]. Similarly, in vestibular nucleus neurons, the fastest firing neurons express Kv3 channels, predominantly to repolarize the neuronal membrane during an AP train. However, in slower-firing neurons, BK channels are responsible for repolarization [180]. This evidence supports the idea that the functional role of BK channels cannot be defined in isolation but critically depends on the context of the other conductances in the cell, as well as their precise localization and composition.
As for principal cells, BK channels also have been reported to control excitability of GABAergic interneurons, but in a lesser extension. For example, in the cortex, action potential repolarization is governed by BK channels in somatostatin (SST)- but not Parvalbumin (PV)-positive interneurons [166]. Interestingly, in the striatum, SST enhances BK currents [146]. If SST released from SST-positive interneurons could act in an autocrine manner to regulate BK channels and control their excitability remains unknown. It is noteworthy, however, that in isolated retinal bipolar cells, SST inhibits BK channels [181]. BK channels also regulate action potential repolarization in striatal cholinergic interneurons [21], regulate firing properties of GABAergic vestibular nucleus neurons [182], and contribute to fAHP in cholinergic neurons of the pontine laterodorsal tegmentum [183]. In the retina, BK channels suppress activation of Ca2+ channels to regulate GABAergic A17 amacrine cell signaling [22]. On the other hand, in fast-spiking stellate cerebellar interneurons, blocking BK channels increases action potential width and Ca2+ influx [184], further supporting a role for BK in regulating GABAergic interneurons function.
BK channels regulate dendritic excitability: In addition to controlling somatic excitability, increasing evidence indicates that BK channels are also present in dendritic compartments [5], where they exert an additional level of control by modulating dendritic excitability. For example, in the somatosensory cortex, BK channels are present in apical dendrites [154], exerting a suppressive effect on the Ca2+ channel to regulate Ca2+ spikes and dendritic excitability [18,144,185]. This control is activity-dependent, as a single action potential or a train of low-frequency action potentials could not activate BK channels [156]. In contrast, a high-frequency train of action potentials can activate BK channels [185]. BK channels also restrict dendritic Ca2+ spikes in CA1 pyramidal cells, thereby generating an inhibitory effect over synaptic potentiation [186] although dendritic BK channels are not activated by backpropagating action potentials [158,187]. Inhibition of BK channels was also shown to improve dendritic Ca2+ spikes in cerebellar Purkinje neurons, thereby facilitating short-term synaptic plasticity mediated by endocannabinoids at distal spines [171] or signal propagation from dendrites to the soma [127]. BK channels also control the magnitude of excitatory potentials in cerebellar Purkinje dendrites [172]. BK-mediated effects for dendritic spike generation could depend on the spine size where they are expressed. For example, in basal dendrites of pyramidal neurons of layer V in the visual cortex, BK channels are activated only in small-head spines, where they suppress excitatory postsynaptic potentials [25]. This effect requires a BK–NMDA receptor complex rather than coupled BK–Cav channels, as Ca2+ influx through NMDAR is required to activate BK channels [25]. A similar coupling between NMDAR and BK channels was also observed in basal dendrites of layer V pyramidal neurons in the barrel cortex, where they avoid the induction of spike-timing-dependent plasticity [188]. However, the BK–NMDAR complex is not observed in all basal dendrites from cortical neurons, where coupling between SK channels and NMDARs to regulate dendritic excitability has also been described [185]. Interestingly, BK-mediated regulation of dendritic excitability in layer 5 neocortical neurons is impaired in mice lacking FMRP [151], thereby suggesting that BK channels might play a role in the pathophysiology of mental deficiency.
BK channels regulate synaptic transmission: BK channels also appear to be essential regulators of neurotransmitter release when expressed at synaptic terminals within different brain circuits [14,16]. Their localization in the active zone appears to be regulated by different synaptic proteins, including RIM-binding proteins [189], synapsin [190], α-catulin [191], and dystrobrevin [191,192]. In the brain, the function of presynaptic BK channels in regulating neurotransmitter release may vary from synapse to synapse [23,24]. For example, BK channels regulate excitatory glutamate-release probability, an effect associated with broadening a presynaptic action potential at CA3–CA3 connections [24]. In contrast, at Schaffer collaterals to CA1 synapses, BK channels contribute to transmitter release only after action potential broadening with 4-aminopyridine, and thus, it has been suggested that BK channels act as ‘emergency brakes’ to exert a protective effect against synaptic hyperactivity [23]. A similar effect was observed at glutamatergic mossy fiber boutons in rat hippocampal slices, where the contribution of BK channels to presynaptic action potential repolarization was not observed under basal conditions due to the presence of a subtype of fast-activating Kv3 channels [164]. Moreover, evidence suggests that postsynaptic BK channels in mature dentate granule cells [27] and thalamocortical circuits [112], could also regulate synaptic transmission via NMDAR-mediated channel activation. Interestingly, NMDARs could be localized at some presynaptic terminals, where their activation regulates transmitter release and plasticity [193,194,195]. Whether a similar BK–NMDA complex is present at presynaptic terminals, thereby regulating transmitter release, remains unknown.
BK-mediated control of synaptic transmission also has been described in rod photoreceptors in the salamander retina [29], and in the brainstem, where they regulate paired-pulse ratio and depression induced by trains of high-frequency stimulation [129]. Although a preferential control exerted by BK channels on excitatory versus inhibitory synapses has been suggested [196,197], it has become clear that BK channels could also control GABA release in different neuronal circuits. For example, in the retina [22] BK channels contribute to GABA release from A17 amacrine cells. In the cortex BK control GABA release from fast-spiking interneurons to II/III layer pyramidal neurons [21] and at GABAergic synapses in the central amygdala [12]. Likewise, BK channels regulate inhibitory glycine release in the spinal cord [198] and acetylcholine release from efferent terminals from the medial olivocochlear system to inner hair cells [199]. These results suggest that by regulating excitatory and inhibitory transmitter release, BK channels might contribute to the excitatory–inhibitory balance necessary for normal brain function. Consistent with this idea, it has been observed that cereblon, a protein related to intellectual disability, downregulates BK channel activity at hippocampal CA3 terminals affecting paired-pulse ratio and short-term plasticity, but not long-term plasticity [200].
Glial BK channels: In addition to being present in neurons, BK channels are also localized in astrocytes [201], where their activation reportedly modulates blood pressure [202,203] and generates vasodilation [204]. Interestingly, astrocytic BK channels mainly express the β4 subunit during neonatal development [12,205], thereby regulating voltage gradients and K+ homeostasis [12], but their contribution to regulating synaptic function remains unclear. In glioblastoma, the most invasive and aggressive primary brain tumor, astrocytic BK channels are likely upregulated to induce water loss and reduce cellular volume [206]. More recently, an interaction between BK channels and the GABA transporter type 3 (GAT3) has also been reported, suggesting that BK channels could regulate GABA homeostasis and signaling by regulating astrocyte function [207]. However, the role of astrocyte BK channels in regulating glutamatergic and GABAergic synaptic function remains unknown.
BK channels are also localized in the spinal cord [208] and cortical microglia [209]. In the spinal cord, BK can induce microglial hyperactivity related to neuropathic pain [210] and potentiate spinal synapses in lamina I spinal neurons [8]. In the brain, evidence suggests that microglial BK channels might participate in BDNF release and induce microglial phagocytosis [209]. BK channels have also been identified in microglial cells of the hippocampus [211,212], which, unlike neuronal BK, might not be developmentally regulated [213]. Whether microglial BK channels may contribute to regulate neuronal excitability and synaptic transmission within selective brain circuits remains unknown.
Lastly, BK channels also seem to be present in oligodendrocyte precursors [214,215] and to a lesser extent in mature oligodendrocytes [214]. In multiple sclerosis lesions, the β4-subunit coding gene KCNMB4, is upregulated in oligodendrocytes [216], but its role is largely unknown.
Could BK channels be activated by endogenous ligands to regulate neuronal function? In addition to being activated by voltage and Ca2+, increasing evidence suggests that different molecules can directly bind to BK channels to regulate their function. For instance, 17β-estradiol activates BK channels via a process that requires the presence of the β1 subunit [217]. Moreover, β4 confers BK channel sensitivity to corticosterone [218]. While evidence suggests that estradiol could regulate hippocampal excitability [219] and regulate glutamatergic synaptic transmission and plasticity in the hippocampus [220] and primary auditory cortex [221], whether these effects are directly mediated by targeting BK channels remains unclear.
Different lipid ligands interact directly with BK channels. For example, the omega-3 fatty acid DHA activates β1 and β4, with minimal effect on β2-containing BK channels [222]. Similarly, arachidonic acid (AA) increases BK channel currents in the presence of β2 subunits [223] and AA metabolites produced by a COX-mediated pathway are responsible for BK channel activation in microglia [8]. In vascular tissues, lipoxygenase (LOX)-metabolites have also been shown to activate BK [224]. In particular, BK channels from coronary artery smooth muscle, which contains β1 and β4 subunits, are activated by AA-lipoxygenase metabolites [225].
The endocannabinoid anandamide and its stable analog, methanandamide, also activate BK currents [226,227]. Moreover, a direct interaction between cannabinoid signaling and the BK channel appears to be involved in the effect of cannabinoids suppressing peripherical firing in injured fibers [228]. However, it is unknown whether endocannabinoids can regulate brain function by targeting BK channels.

7. BK Channel Dysfunction in Neurological Disorders

Given the critical role of BK channels in regulating neuronal excitability and synaptic function, it is not surprising that dysfunction of BK channels has been implicated in several neurological disorders. These include epilepsy, fragile X syndrome, mental retardation, autism, movement disorders, and chronic pain [229]. The use of BK channels in animal models of BK loss-of-function (LOF) and gain-of-function (GOF) mutants has increased our current understanding of the role of BK channels in the etiology of these disturbances. For example, GOF mutations are more common than LOF alleles for patients with paroxysmal non-kinesigenic dyskinesia (PNKD), while non-PKND movement disorders were observed in LOF mutations [230]. In a GOF knock-in murine model carrying a human BK-D434G mutation, mice showed increased susceptibility to generalized seizures and motor deficits related to hyperexcitability in cortical pyramidal and cerebellar Purkinje neurons. Importantly, blocking BK channels with paxilline suppressed seizures, motor disturbances, and neuronal hyperexcitability [155]. Similarly, a D434G mutation that produces an increase in Ca2+ sensitivity that potentiates BK channel activity [231] was described in individuals affected by epileptic seizures and/or PNKD [232]. Moreover, a rare disease called KCNMA1-linked channelopathy, is also characterized by seizures and abnormal movements associated with the clinical detection of a KCNMA1 variant [232]. Another mutation, the N995S variant, is present in individuals with epilepsy but not PNKD. Such a mutation in the RCK2 domain increases BK channel currents by increasing voltage sensitivity but does not affect Ca2+-dependent activation. In BK-N995S/β4 channels, the effect of the mutation is slightly more pronounced [233]. However, in a mouse model with the N995S mutation, PNKD-like symptoms could appear after stress exposure [232].
We recall here that targeted deletion of the KCNMB4 gene coding for β4 subunits narrows action potentials and increases firing in dentate granule cells of the hippocampus, possibly leading to an enhancement of transmitter release and temporal lobe seizures [72]. Consistent with the involvement of the β4 subunit in epilepsy, pilocarpine-induced seizures generated a decrease in β4 subunits in dentate gyrus granular cells and a reduction in type II BK channels [234]. While β4 subunits seem to play an important role in epilepsy, a mutation in the gene encoding the accessory β3 subunits (KCNMB3b β3b-truncation) have also been frequently observed in patients affected by idiopathic generalized epilepsy [235].
Functional alterations of BK channel expression are also linked to neuropathic pain. A downregulation of BK channels in dorsal root ganglions and the spinal dorsal horn is accompanied by neuropathic pain [236,237,238]. Nerve injury promotes an increase in N-type Ca2+ channels in the spinal dorsal horn mediated by an upregulation of Cav2δ subunits, increasing pain signaling. Expressing the BK channel N-terminus, a domain that binds the Cav2δ subunit, promotes a reduction in the expression of N-type Ca2+ channels and analgesia [239]. Additionally, the pharmacological blockade of spinal BK channels with iberiotoxin produces mechanical hyperalgesia. In contrast, activation of the BK channel reduces tactile allodynia and mechanical and thermal hyperalgesia, in nerve-ligated rats [236]. Similarly, decreasing spinal β3 subunit expression reportedly attenuated neuropathic pain produced by nerve injury in mice [240], further supporting a role for BK channels in regulating pain.
BK channels are also involved in motor and cognitive diseases. For instance, in homozygous mice lacking α subunits, muscular tremors, muscular weakness, and gait disturbances have been reported. However, in an open field test, these animal models showed normal distance for locomotor behavior. In addition, these mice also display deficits in learning, with a delay in acquiring a new task but not in memory once the task was learned [241]. Similarly, a mutation (BKG354S) in the α subunit has recently been reported in a child with congenital and progressive cerebellar ataxia with cognitive impairment. This mutation dramatically reduced single-channel conductance and ion selectivity and reduced neurite outgrowth, cell viability, and mitochondrial content [242].
Besides humans, GOF mutations are also present in Drosophila. There, the SLO E366G mutation (equivalent to the human mutation D434G) generates motor disturbances in both larvae and adult stages [243]. In C.elegans, deleting the Shank gene Shn-1, a synaptic scaffolding protein related to autism, produces a decrease in BK channel currents and clustering at muscles and neurons [244]. Interestingly, BK channels may play a role in the physiopathology of fragile X syndrome [245] as accessory β4 subunits interact with FMRP proteins, whose loss causes the fragile X syndrome, characterized by intellectual disabilities associated with language deficits, hyperactivity, autistic behavior, and seizures. In the absence of β4 subunits, FMRP decrease channel activation and deactivation rates, but in the presence of β4, FMRP increase channel opening [81]. BK channel agonists have a major impact on rescuing social and cognitive impairments [246]. However, the diversity of BK channel expression within different neuronal populations and brain circuits and the poor selectivity of drugs, suggest that one should take the use of these compounds in clinical practice with caution.
A role for BK channels in modulating learning and memory processes also has been suggested [160,247,248,249]. For example, if a methyl-donor-rich diet is supplied to mice before mating, the offspring show disturbances in hippocampus-related memory tasks and long-term potentiation of synaptic transmission related to decreased β2 subunit expression in the hippocampus [250]. Additionally, anxiety and cognitive disturbances induced by thalidomide were correlated to increased BK channel activity. These alterations can be rescued by the BK channel blocker paxilline [251]. Similarly, BK channel currents are reduced in pyramidal neurons of the anterior cingulate cortex in a mechanical allodynia model generating anxiety behaviors in mice. Upregulated expression of β4 subunits in presynaptic and postsynaptic domains and increased neuronal excitability and presynaptic neurotransmitter release accompanied this phenomenon. BK channel activation with NS1619 reversed anxiety-like behaviors and the mechanical pain threshold [73]. Last but not least, a genome-wide association study reported that a single-nucleotide polymorphism (SNP) in the gene encoding α subunits (KCNMA1, rs16934131) was associated with late-onset Alzheimer’s disease (AD) risk in humans [252].
On the other hand, an SNP in the gene encoding β2 (KCNMB2, rs9637454) was strongly associated with hippocampal sclerosis, a comorbid neuropathological feature of AD [253]. While the functional consequence of these point mutations is unknown, rodent AD models have shed some light on how BK channels may contribute to AD pathogenesis. For example, in TgCRND8 mice; a mouse model for AD which overexpresses the amyloid precursor protein due to containing a double human amyloid precursor protein mutation; a decreased basal transmission associated with a reduced decay of afferent volleys has been reported. This difference was recovered when blockers for BK channels were present [254]. Moreover, intracellular injection of amyloid-β1–42 (Aβ1–42) peptides into neocortical pyramidal cells suppresses BK channel activity, which broadens evoked AP spike width and indirectly enhances Ca2+ influx [255]. Similarly, Ab42 oligomers, stimulate spontaneous transient Ca2+ release from RyRs. RyRs activate presynaptic BK channels and reduce AMPA-mediated neurotransmission in hippocampal neurons [256]. Conversely, in the triple transgenic AD mouse model 3xTg, increased basal transmission and pyramidal excitability and decreased synaptic plasticity at the hippocampal CA1 area were related to a hypofunction of BK channels. 3xTg also displayed poor performance in non-spatial and spatial memory tasks, a dysfunction that was improved when a BK channel activator was delivered through ventricular injection [257]. While this evidence suggests a potential role of BK channels in the pathophysiology of AD, a direct link between Aβ42 oligomers and BK channel activation at the early stages of AD remains to be identified.

8. Conclusions and Future Directions

Increasing evidence indicates that, in addition to its well-known role in regulating neuronal excitability, BK channels also regulate dendritic excitability and synaptic transmission in different brain circuits. Their presynaptic and postsynaptic expression within a specific synapse suggests they are also well positioned to regulate synaptic function of excitatory and inhibitory neurons. This effect may vary from synapse to synapse (i.e., either increase or decrease) and might also regulate synaptic plasticity. While indirect evidence suggests that BK channels can regulate long-term potentiation and depression, whether BK channels play a pivotal role in the induction or expression of a different form of synaptic plasticity needs further investigation. Similarly, their expression in glial cells has yet to be entirely understood. For instance, the role of BK channels regulating gliotransmission, thereby modulating synaptic brain function, remains largely unexplored. Future work will expand our understanding of the role of glial BK channels, particularly their contribution to regulating GABA homeostasis and signaling, as they have been reported to directly interact with the GABA transporter in astrocytes [207]. Notably, lipid molecules, including AA derivates and endocannabinoids, are known to stimulate BK channels in heterologous and peripheral systems. However, whether these lipids regulate brain function and behavior in a BK-dependent manner remains unknown. Given the ubiquitous expression of BK channels throughout the brain, it is not surprising that they have been implicated in several neurological disorders, such as epilepsy, mental retardation, autism, AD, anxiety, movement disorders, and chronic pain [229]. Using LOF or GOF animal models has increased our current understanding of the role of BK channels in the etiology of these disturbances. However, further work is required for better development of novel therapeutic strategies and interventions that may ameliorate these devastating disorders.

Author Contributions

C.A.-G., I.S. and R.A.-S., original draft preparation, Tables and Figures. A.E.C. and R.L. guided the research and wrote the paper. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Chilean government through FONDECYT Regular #1201848 (A.E.C.), #1190203 (R.L.), and by ANID Millennium Science Initiative Program (ACE210014 to A.E.C. and R.L.). C.A.G. and R.A.S. were supported by a PhD fellowship from ANID #21201603 (C.A.G.) and #21221040 (R.A.S.). We apologize to those authors whose work we could not cite due to space limitations.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Pyott, S.; Duncan, R. BK Channels in the Vertebrate Inner Ear. Int. Rev. Neurobiol. 2016, 128, 369–399. [Google Scholar] [CrossRef] [PubMed]
  2. Krishnamoorthy-Natarajan, G.; Koide, M. BK Channels in the Vascular System. Int. Rev. Neurobiol. 2016, 128, 401–438. [Google Scholar] [CrossRef] [PubMed]
  3. Harvey, J.R.M.; Plante, A.E.; Meredith, A.L. Ion Channels Controlling Circadian Rhythms in Suprachiasmatic Nucleus Excitability. Physiol. Rev. 2020, 100, 1415–1454. [Google Scholar] [CrossRef]
  4. Knaus, H.; Schwarzer, C.; Koch, R.; Eberhart, A.; Kaczorowski, G.; Glossmann, H.; Wunder, F.; Pongs, O.; Garcia, M.; Sperk, G. Distribution of high-conductance Ca(2+)-activated K+ channels in rat brain: Targeting to axons and nerve terminals. J. Neurosci. 1996, 16, 955–963. [Google Scholar] [CrossRef] [PubMed]
  5. Sausbier, U.; Sausbier, M.; Sailer, C.A.; Arntz, C.; Knaus, H.-G.; Neuhuber, W.; Ruth, P. Ca2+-activated K+ channels of the BK-type in the mouse brain. Histochem. Cell Biol. 2006, 125, 725–741. [Google Scholar] [CrossRef]
  6. Wanner, S.G.; Koch, R.O.; Koschak, A.; Trieb, M.; Garcia, M.L.; Kaczorowski, G.J.; Knaus, H.-G. High-Conductance Calcium-Activated Potassium Channels in Rat Brain: Pharmacology, Distribution, and Subunit Composition. Biochemistry 1999, 38, 5392–5400. [Google Scholar] [CrossRef]
  7. González-Sanabria, N.; Echeverría, F.; Segura, I.; Alvarado-Sánchez, R.; Latorre, R. BK in Double-Membrane Organelles: A Biophysical, Pharmacological, and Functional Survey. Front. Physiol. 2021, 12, 761474. [Google Scholar] [CrossRef]
  8. Hayashi, Y.; Morinaga, S.; Zhang, J.; Satoh, Y.; Meredith, A.; Nakata, T.; Wu, Z.; Kohsaka, S.; Inoue, K.; Nakanishi, H. BK channels in microglia are required for morphine-induced hyperalgesia. Nat. Commun. 2016, 7, 11697. [Google Scholar] [CrossRef]
  9. Kyle, B.D.; Braun, A.P. The regulation of BK channel activity by pre- and post-translational modifications. Front. Physiol. 2014, 5, 316. [Google Scholar] [CrossRef]
  10. Latorre, R.; Castillo, K.; Carrasquel-Ursulaez, W.; Sepulveda, R.V.; Gonzalez-Nilo, F.; Gonzalez, C.; Alvarez, O. Molecular Determinants of BK Channel Functional Diversity and Functioning. Physiol. Rev. 2017, 97, 39–87. [Google Scholar] [CrossRef]
  11. Li, B.; Gao, T.-M. Functional Role of Mitochondrial and Nuclear BK Channels. Int. Rev. Neurobiol. 2016, 128, 163–191. [Google Scholar] [CrossRef]
  12. Seidel, K.N.; Derst, C.; Salzmann, M.; Höltje, M.; Priller, J.; Markgraf, R.; Heinemann, S.H.; Heilmann, H.; Skatchkov, S.N.; Eaton, M.J.; et al. Expression of the voltage- and Ca2+-dependent BK potassium channel subunits BKβ1 and BKβ4 in rodent astrocytes. Glia 2011, 59, 893–902. [Google Scholar] [CrossRef]
  13. Seifert, G.; Henneberger, C.; Steinhäuser, C. Diversity of astrocyte potassium channels: An update. Brain Res. Bull. 2018, 136, 26–36. [Google Scholar] [CrossRef]
  14. Contet, C.; Goulding, S.P.; Kuljis, D.A.; Barth, A.L. BK channel sin the CNS. Int. Rev. Neurobiol. 2016, 128, 281–342. [Google Scholar]
  15. Deng, P.-Y.; Rotman, Z.; Blundon, J.A.; Cho, Y.; Cui, J.; Cavalli, V.; Zakharenko, S.; Klyachko, V.A. FMRP Regulates Neurotransmitter Release and Synaptic Information Transmission by Modulating Action Potential Duration via BK Channels. Neuron 2013, 77, 696–711. [Google Scholar] [CrossRef]
  16. Griguoli, M.; Sgritta, M.; Cherubini, E. Presynaptic BK channels control transmitter release: Physiological relevance and potential therapeutic implications. J. Physiol. 2016, 594, 3489–3500. [Google Scholar] [CrossRef]
  17. Sausbier, M.; Hu, H.; Arntz, C.; Feil, S.; Kamm, S.; Adelsberger, H.; Sausbier, U.; Sailer, C.A.; Feil, R.; Hofmann, F.; et al. Cerebellar ataxia and Purkinje cell dysfunction caused by Ca2+-activated K+channel deficiency. Proc. Natl. Acad. Sci. USA 2004, 101, 9474–9478. [Google Scholar] [CrossRef]
  18. Bock, T.; Stuart, G.J. The Impact of BK Channels on Cellular Excitability Depends on their Subcellular Location. Front. Cell. Neurosci. 2016, 10, 206. [Google Scholar] [CrossRef]
  19. Storm, J.F. Understanding the Brain Through the Hippocampus the Hippocampal Region as a Model for Studying Brain Structure and Function. Prog. Brain Res. 1990, 83, 161–187. [Google Scholar]
  20. Bielefeldt, K.; Jackson, M.B. A calcium-activated potassium channel causes frequency-dependent action-potential failures in a mammalian nerve terminal. J. Neurophysiol. 1993, 70, 284–298. [Google Scholar] [CrossRef]
  21. Goldberg, J.A.; Wilson, C.J. Control of spontaneous firing patterns by the selective coupling of calcium currents to calcium-activated potassium currents in striatal cholinergic interneurons. J. Neurosci. 2005, 25, 10230–10238. [Google Scholar] [CrossRef]
  22. Grimes, W.N.; Li, W.; Chávez, A.E.; Diamond, J.S. BK channels modulate pre- and postsynaptic signaling at reciprocal synapses in retina. Nat. Neurosci. 2009, 12, 585–592. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Hu, H.; Shao, L.-R.; Chavoshy, S.; Gu, N.; Trieb, M.; Behrens, R.; Laake, P.; Pongs, O.; Knaus, H.G.; Ottersen, O.P.; et al. Presynaptic Ca2+-Activated K+Channels in Glutamatergic Hippocampal Terminals and Their Role in Spike Repolarization and Regulation of Transmitter Release. J. Neurosci. 2001, 21, 9585–9597. [Google Scholar] [CrossRef] [PubMed]
  24. Raffaelli, G.; Saviane, C.; Mohajerani, M.H.; Pedarzani, P.; Cherubini, E. BK potassium channels control transmitter release at CA3-CA3 synapses in the rat hippocampus. J. Physiol. 2004, 557, 147–157. [Google Scholar] [CrossRef]
  25. Tazerart, S.; Blanchard, M.G.; Miranda-Rottmann, S.; Mitchell, D.E.; Pina, B.N.; Thomas, C.I.; Kamasawa, N.; Araya, R. Selective activation of BK channels in small-headed dendritic spines suppresses excitatory postsynaptic potentials. J. Physiol. 2022, 600, 2165–2187. [Google Scholar] [CrossRef] [PubMed]
  26. Guo, Y.-Y.; Liu, S.-B.; Cui, G.-B.; Ma, L.; Feng, B.; Xing, J.-H.; Yang, Q.; Li, X.-Q.; Wu, Y.-M.; Xiong, L.-Z.; et al. Acute stress induces down-regulation of large-conductance Ca2+-activated potassium channels in the lateral amygdala. J. Physiol. 2012, 590, 875–886. [Google Scholar] [CrossRef]
  27. Zhang, J.; Guan, X.; Li, Q.; Meredith, A.L.; Pan, H.-L.; Yan, J. Glutamate-activated BK channel complexes formed with NMDA receptors. Proc. Natl. Acad. Sci. USA 2018, 115, E9006–E9014. [Google Scholar] [CrossRef]
  28. Cavaccini, A.; Gritti, M.; Giorgi, A.; Locarno, A.; Heck, N.; Migliarini, S.; Bertero, A.; Mereu, M.; Margiani, G.; Trusel, M.; et al. Serotonergic Signaling Controls Input-Specific Synaptic Plasticity at Striatal Circuits. Neuron 2018, 98, 801–816.e7. [Google Scholar] [CrossRef]
  29. Xu, J.W.; Slaughter, M.M. Large-Conductance Calcium-Activated Potassium Channels Facilitate Transmitter Release in Salamander Rod Synapse. J. Neurosci. 2005, 25, 7660–7668. [Google Scholar] [CrossRef]
  30. Eisenman, G.; Latorre, R.; Miller, C. Multi-ion conduction and selectivity in the high-conductance Ca++-activated K+ channel from skeletal muscle. Biophys. J. 1986, 50, 1025–1034. [Google Scholar] [CrossRef]
  31. Latorre, R.; Miller, C. Conduction and selectivity in potassium channels. J. Membr. Biol. 1983, 71, 11–30. [Google Scholar] [CrossRef]
  32. Yellen, G. Ionic permeation and blockade in Ca2+-activated K+ channels of bovine chromaffin cells. J. Gen. Physiol. 1984, 84, 157–186. [Google Scholar] [CrossRef]
  33. Adelman, J.P.; Shen, K.-Z.; Kavanaugh, M.P.; Warren, R.A.; Wu, Y.-N.; Lagrutta, A.; Bond, C.T.; North, R.A. Calcium-activated potassium channels expressed from cloned complementary DNAs. Neuron 1992, 9, 209–216. [Google Scholar] [CrossRef]
  34. Butler, A.; Tsunoda, S.; McCobb, D.P.; Wei, A.; Salkoff, L. mSlo, a complex mouse gene encoding “maxi” calcium-activated potassium channels. Science 1993, 261, 221–224. [Google Scholar] [CrossRef]
  35. Meera, P.; Wallner, M.; Song, M.; Toro, L. Large conductance voltage- and calcium-dependent K+ channel, a distinct member of voltage-dependent ion channels with seven N-terminal transmembrane segments (SO-S6), an extracellular N terminus, and an intracellular (S9-S10) C terminus. Proc. Natl. Acad. Sci. USA 1997, 94, 14066–14071. [Google Scholar] [CrossRef]
  36. Bao, L.; Rapin, A.M.; Holmstrand, E.C.; Cox, D.H. Elimination of the BKCa Channel’s High-Affinity Ca2+ Sensitivity. J. Gen. Physiol. 2002, 120, 173–189. [Google Scholar] [CrossRef]
  37. Salkoff, L.; Butler, A.; Ferreira, G.; Santi, C.; Wei, A. High-conductance potassium channels of the SLO family. Nat. Rev. Neurosci. 2006, 7, 921–931. [Google Scholar] [CrossRef]
  38. Tao, X.; Hite, R.K.; MacKinnon, R. Cryo-EM structure of the open high-conductance Ca2+-activated K+ channel. Nature 2016, 541, 46–51. [Google Scholar] [CrossRef]
  39. Xia, X.-M.; Zeng, X.; Lingle, C.J. Multiple regulatory sites in large-conductance calcium-activated potassium channels. Nature 2002, 418, 880–884. [Google Scholar] [CrossRef]
  40. McManus, O.B.; Blatz, A.L.; Magleby, K.L. Inverse relationship of the durations of adjacent open and shut intervals for Cl and K channels. Nature 1985, 317, 625–627. [Google Scholar] [CrossRef]
  41. Pallotta, B.S. Calcium-activated potassium channels in rat muscle inactivate from a short-duration open state. J. Physiol. 1985, 363, 501–516. [Google Scholar] [CrossRef] [PubMed]
  42. Rothberg, B.S.; Magleby, K.L. Voltage and Ca2+ Activation of Single Large-Conductance Ca2+-Activated K+ Channels Described by a Two-Tiered Allosteric Gating Mechanism. J. Gen. Physiol. 2000, 116, 75–100. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Horrigan, F.T.; Aldrich, R.W. Coupling between Voltage Sensor Activation, Ca2+ Binding and Channel Opening in Large Conductance (BK) Potassium Channels. J. Gen. Physiol. 2002, 120, 267–305. [Google Scholar] [CrossRef] [PubMed]
  44. Carrasquel-Ursulaez, W.; Contreras, G.F.; Sepúlveda, R.V.; Aguayo, D.; González-Nilo, F.; González, C.; Latorre, R. Hydrophobic interaction between contiguous residues in the S6 transmembrane segment acts as a stimuli integration node in the BK channel. J. Gen. Physiol. 2014, 145, 61–74. [Google Scholar] [CrossRef]
  45. Lorenzo-Ceballos, Y.; Carrasquel-Ursulaez, W.; Castillo, K.; Alvarez, O.; Latorre, R. Calcium-driven regulation of voltagesensingdomains in BK channels. Elife 2019, 8, 1–24. [Google Scholar] [CrossRef]
  46. Hite, R.K.; Tao, X.; MacKinnon, R. Structural basis for gating the high-conductance Ca2+-activated K+ channel. Nature 2016, 541, 52–57. [Google Scholar] [CrossRef]
  47. Ma, Z.; Lou, X.J.; Horrigan, F.T. Role of Charged Residues in the S1–S4 Voltage Sensor of BK Channels. J. Gen. Physiol. 2006, 127, 309–328. [Google Scholar] [CrossRef]
  48. Carrasquel-Ursulaez, W.; Segura, I.; Díaz-Franulic, I.; Echeverría, F.; Lorenzo-Ceballos, Y.; Espinoza, N.; Rojas, M.; Garate, J.A.; Perozo, E.; Alvarez, O. Mechanism of voltage sensing in Ca2+- and voltage-activated K+ (BK) channels. Proc. Natl. Acad. Sci. USA 2022, 119, 1–11. [Google Scholar] [CrossRef]
  49. Zeng, X.-H.; Xia, X.-M.; Lingle, C.J. Divalent Cation Sensitivity of BK Channel Activation Supports the Existence of Three Distinct Binding Sites. J. Gen. Physiol. 2005, 125, 273–286. [Google Scholar] [CrossRef]
  50. Zhou, Y.; Zeng, X.-H.; Lingle, C.J. Barium ions selectively activate BK channels via the Ca 2+ -bowl site. Proc. Natl. Acad. Sci. USA 2012, 109, 11413–11418. [Google Scholar] [CrossRef]
  51. Lee, C.-H.; MacKinnon, R. Structures of the Human HCN1 Hyperpolarization-Activated Channel. Cell 2017, 168, 111–120.e11. [Google Scholar] [CrossRef]
  52. Wang, W.; MacKinnon, R. Cryo-EM Structure of the Open Human Ether-à-go-go -Related K + Channel hERG. Cell 2017, 169, 422–430.e10. [Google Scholar] [CrossRef] [Green Version]
  53. Tang, Q.-Y.; Zeng, X.-H.; Lingle, C.J. Closed-channel block of BK potassium channels by bbTBA requires partial activation. J. Gen. Physiol. 2009, 134, 409–436. [Google Scholar] [CrossRef]
  54. Thompson, J.; Begenisich, T. Selectivity filter gating in large-conductance Ca2+-activated K+ channels. J. Gen. Physiol. 2012, 139, 235–244. [Google Scholar] [CrossRef]
  55. Wilkens, C.M.; Aldrich, R.W. State-independent Block of BK Channels by an Intracellular Quaternary Ammonium. J. Gen. Physiol. 2006, 128, 347–364. [Google Scholar] [CrossRef]
  56. Jia, Z.; Yazdani, M.; Zhang, G.; Cui, J.; Chen, J. Hydrophobic gating in BK channels. Nat. Commun. 2018, 9, 3408. [Google Scholar] [CrossRef]
  57. Torres, Y.P.; Granados, S.T.; Latorre, R. Pharmacological consequences of the coexpression of BK channel α and auxiliary β subunits. Front Physiol. 2014, 5, 383. [Google Scholar] [CrossRef]
  58. Orio, P.; Rojas, P.; Ferreira, G.; Latorre, R. New Disguises for an Old Channel: MaxiK Channel β-Subunits. Physiology 2002, 17, 156–161. [Google Scholar] [CrossRef]
  59. Zhou, Y.; Lingle, C.J. Paxilline inhibits BK channels by an almost exclusively closed-channel block mechanism. J. Gen. Physiol. 2014, 144, 415–440. [Google Scholar] [CrossRef]
  60. Meera, P.; Wallner, M.; Toro, L. A neuronal β subunit (KCNMB4) makes the large conductance, voltage- and Ca2+ -activated K + channel resistant to charybdotoxin and iberiotoxin. Proc. Natl. Acad. Sci. USA 2000, 97, 5562–5567. [Google Scholar] [CrossRef]
  61. Ha, T.S.; Heo, M.-S.; Park, C.-S. Functional Effects of Auxiliary β4-Subunit on Rat Large-Conductance Ca2+-Activated K+ Channel. Biophys. J. 2004, 86, 2871–2882. [Google Scholar] [CrossRef] [PubMed]
  62. Wang, B.; Rothberg, B.S.; Brenner, R. Mechanism of β4 Subunit Modulation of BK Channels. J. Gen. Physiol. 2006, 127, 449–465. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Jaffe, D.; Wang, B.; Brenner, R. Shaping of action potentials by type I and type II large-conductance Ca2+-activated K+ channels. Neuroscience 2011, 192, 205–218. [Google Scholar] [CrossRef] [PubMed]
  64. Wang, B.; Jaffe, D.B.; Brenner, R. Current understanding of iberiotoxin-resistant BK channels in the nervous system. Front. Physiol. 2014, 5, 382. [Google Scholar] [CrossRef]
  65. Reinhart, P.H.; Chung, S.; Levitan, I.B. A family of calcium-dependent potassium channels from rat brain. Neuron 1989, 2, 1031–1041. [Google Scholar] [CrossRef]
  66. Behrens, R.; Nolting, A.; Reimann, F.; Schwarz, M.; Waldschütz, R.; Pongs, O. hKCNMB3 and hKCNMB4, cloning and characterization of two members of the large-conductance calcium-activated potassium channel β subunit family. FEBS Lett. 2000, 474, 99–106. [Google Scholar] [CrossRef]
  67. Lippiat, J.D.; Standen, N.B.; Harrow, I.D.; Phillips, S.C.; Davies, N.W. Properties of BKCa channels formed by bicistronic expression of hSloα and β1-4 subunits in HEK293 cells. J. Membr. Biol. 2003, 192, 141–148. [Google Scholar]
  68. Gan, G.; Yi, H.; Chen, M.; Sun, L.; Li, W.; Wu, Y.; Ding, J. Structural Basis for Toxin Resistance of β4-Associated Calcium-activated Potassium (BK) Channels. J. Biol. Chem. 2008, 283, 24177–24184. [Google Scholar] [CrossRef]
  69. Tao, X.; Mackinnon, R. Molecular structures of the human slo1 k+ channel in complex with b4. Elife 2019, 8, e51409. [Google Scholar] [CrossRef]
  70. Candia, S.; Garcia, M.; Latorre, R. Mode of action of iberiotoxin, a potent blocker of the large conductance Ca(2+)-activated K+ channel. Biophys. J. 1992, 63, 583–590. [Google Scholar] [CrossRef]
  71. Weiger, T.M.; Holmqvist, M.H.; Levitan, I.B.; Clark, F.T.; Sprague, S.; Huang, W.-J.; Ge, P.; Wang, C.; Lawson, D.; Jurman, M.E.; et al. A Novel Nervous System β Subunit that Downregulates Human Large Conductance Calcium-Dependent Potassium Channels. J. Neurosci. 2000, 20, 3563–3570. [Google Scholar] [CrossRef]
  72. Brenner, R.; Chen, Q.H.; Vilaythong, A.; Toney, G.M.; Noebels, J.; Aldrich, R.W. BK channel β4 subunit reduces dentate gyrus excitability and protects against temporal lobe seizures. Nat. Neurosci. 2005, 8, 1752–1759. [Google Scholar] [CrossRef]
  73. Zhao, H.; Xue, Q.; Li, C.; Wang, Q.; Han, S.; Zhou, Y.; Yang, T.; Xie, Y.; Fu, H.; Lu, C.; et al. Upregulation of Beta4 subunit of BKCa channels in the anterior cingulate cortex contributes to mechanical allodynia associated anxiety-like behaviors. Mol. Brain 2020, 13, 1–18. [Google Scholar] [CrossRef] [Green Version]
  74. Piwonska, M.; Wilczek, E.; Szewczyk, A.; Wilczynski, G. Differential distribution of Ca2+-activated potassium channel β4 subunit in rat brain: Immunolocalization in neuronal mitochondria. Neuroscience 2008, 153, 446–460. [Google Scholar] [CrossRef]
  75. Benton, M.D.; Lewis, A.H.; Bant, J.; Raman, I.M. Iberiotoxin-sensitive and -insensitive BK currents in Purkinje neuron somata. J. Neurophysiol. 2013, 109, 2528–2541. [Google Scholar] [CrossRef]
  76. Jiang, Z.; Wallner, M.; Meera, P.; Toro, L. Human and Rodent MaxiK Channel β-Subunit Genes: Cloning and Characterization. Genomics 1999, 55, 57–67. [Google Scholar] [CrossRef]
  77. Salzmann, M.; Seidel, K.N.; Bernard, R.; Prüss, H.; Veh, R.W.; Derst, C. BKβ1 Subunits Contribute to BK Channel Diversity in Rat Hypothalamic Neurons. Cell. Mol. Neurobiol. 2010, 30, 967–976. [Google Scholar] [CrossRef]
  78. Tseng-Crank, J.; Godinot, N.; Johansen, T.E.; Ahring, P.K.; Strøbæk, D.; Mertz, R.; Foster, C.D.; Olesen, S.P.; Reinhart, P.H. Cloning, expression, and distribution of a Ca2+-activated K+ channel β-subunit from human brain. Proc. Natl. Acad. Sci. USA 1996, 93, 9200–9205. [Google Scholar] [CrossRef]
  79. Wynne, P.M.; Puig, S.I.; Martin, G.E.; Treistman, S.N. Compartmentalized β subunit distribution determines characteristics and ethanol sensitivity of somatic, dendritic, and terminal large-conductance calcium-activated potassium channels in the rat central nervous system. J. Pharmacol. Exp. Ther. 2009, 329, 978–986. [Google Scholar] [CrossRef]
  80. Deng, P.-Y.; Klyachko, V.A. Genetic upregulation of BK channel activity normalizes multiple synaptic and circuit defects in a mouse model of fragile X syndrome. J. Physiol. 2015, 594, 83–97. [Google Scholar] [CrossRef]
  81. Kshatri, A.; Cerrada, A.; Gimeno, R.; Bartolomé-Martín, D.; Rojas, P.; Giraldez, T. Differential regulation of BK channels by fragile X mental retardation protein. J. Gen. Physiol. 2020, 152, e201912502. [Google Scholar] [CrossRef] [PubMed]
  82. Pyott, S.; Glowatzki, E.; Trimmer, J.S.; Aldrich, R.W. Extrasynaptic Localization of Inactivating Calcium-Activated Potassium Channels in Mouse Inner Hair Cells. J. Neurosci. 2004, 24, 9469–9474. [Google Scholar] [CrossRef] [PubMed]
  83. Faber, E.S.L.; Sah, P. Ca2+ -activated K+ (BK) channel inactivation contributes to spike broadening during repetitive firing in the rat lateral amygdala. J Physiol. 2003, 552, 483–497. [Google Scholar] [CrossRef] [PubMed]
  84. Hicks, G.A.; Marrion, N.V. Ca2+-dependent inactivation of large conductance Ca2+-activated K+ (BK) channels in rat hippocampal neurones produced by pore block from an associated particle. J. Physiol. 1998, 508, 721–734. [Google Scholar] [CrossRef]
  85. Li, W.; Gao, S.-B.; Lv, C.-X.; Wu, Y.; Guo, Z.-H.; Ding, J.-P.; Xu, T. Characterization of voltage-and Ca2+-activated K+ channels in rat dorsal root ganglion neurons. J. Cell. Physiol. 2007, 212, 348–357. [Google Scholar] [CrossRef]
  86. Whitt, J.P.; Montgomery, J.R.; Meredith, A.L. BK channel inactivation gates daytime excitability in the circadian clock. Nat. Commun. 2016, 7, 10837. [Google Scholar] [CrossRef]
  87. Meredith, A.L.; Wiler, S.W.; Miller, B.H.; Takahashi, J.S.; Fodor, A.A.; Ruby, N.F.; Aldrich, R.W. BK calcium-activated potassium channels regulate circadian behavioral rhythms and pacemaker output. Nat. Neurosci. 2006, 9, 1041–1049. [Google Scholar] [CrossRef]
  88. Pitts, G.R.; Ohta, H.; McMahon, D.G. Daily rhythmicity of large-conductance Ca2+-activated K+ currents in suprachiasmatic nucleus neurons. Brain Res. 2006, 1071, 54–62. [Google Scholar] [CrossRef]
  89. Berkefeld, H.; Fakler, B.; Schulte, U. Ca2+-activated K+ channels: From protein complexes to function. Physiol. Rev. 2010, 90, 1437–1459. [Google Scholar] [CrossRef]
  90. Shah, K.R.; Guan, X.; Yan, J. Structural and Functional Coupling of Calcium-Activated BK Channels and Calcium-Permeable Channels Within Nanodomain Signaling Complexes. Front. Physiol. 2022, 12, 2556. [Google Scholar] [CrossRef]
  91. Sweet, T.B.; Cox, D.H. Measurements of the BK Ca channel’s high-affinity Ca 2+ binding constants: Effects of membrane voltage. J. Gen. Physiol. 2008, 132, 491–505. [Google Scholar] [CrossRef]
  92. Fakler, B.; Adelman, J.P. Control of KCa Channels by Calcium Nano/Microdomains. Neuron 2008, 59, 873–881. [Google Scholar] [CrossRef]
  93. Naraghi, M.; Neher, E. Linearized Buffered Ca2+Diffusion in Microdomains and Its Implications for Calculation of [Ca2+] at the Mouth of a Calcium Channel. J. Neurosci. 1997, 17, 6961–6973. [Google Scholar] [CrossRef] [Green Version]
  94. Bellono, N.W.; Leitch, D.B.; Julius, D. Molecular basis of ancestral vertebrate electroreception. Nature 2017, 543, 391–396. [Google Scholar] [CrossRef]
  95. Berkefeld, H.; Fakler, B. Repolarizing Responses of BKCa-Cav Complexes Are Distinctly Shaped by Their Cav Subunits. J. Neurosci. 2008, 28, 8238–8245. [Google Scholar] [CrossRef]
  96. Berkefeld, H.; Sailer, C.A.; Bildl, W.; Rohde, V.; Thumfart, J.O.; Eble, S.; Klugbauer, N.; Reisinger, E.; Bischofberger, J.; Oliver, D.; et al. BKCa-Cav channel complexes mediate rapid and localized Ca 2+-activated K+ signaling. Science 2006, 314, 615–620. [Google Scholar] [CrossRef]
  97. Marrion, N.V.; Tavalin, S.J. Selective activation of Ca2+-activated K+ channels by co-localized Ca2+ channels in hippocampal neurons. Nature 1998, 395, 900–905. [Google Scholar] [CrossRef]
  98. Rehak, R.; Bartoletti, T.M.; Engbers, J.D.T.; Berecki, G.; Turner, R.W.; Zamponi, G.W. Low Voltage Activation of KCa1.1 Current by Cav3-KCa1.1 Complexes. PLoS ONE 2013, 8, e61844. [Google Scholar] [CrossRef]
  99. Müller, A.; Kukley, M.; Uebachs, M.; Beck, H.; Dietrich, D. Nanodomains of single Ca2+ channels contribute to action potential repolarization in cortical neurons. J. Neurosci. 2007, 27, 483–495. [Google Scholar] [CrossRef]
  100. Gola, M.; Crest, M. Colocalization of active KCa channels and Ca2+ channels within Ca2+ Domains in helix neurons. Neuron 1993, 10, 689–699. [Google Scholar] [CrossRef]
  101. Monreal, M.R.; Pérez, J.Q.; Escalera, M.P.; Lazalde, A.R.; Bonilla, M.E.P. Development of the L-type CaV/BK Complex Simulator (II): Estimation of the distance between the channels. South Florida J. Dev. 2021, 2, 1258–1265. [Google Scholar] [CrossRef]
  102. Edgerton, J.R.; Reinhart, P.H. Distinct contributions of small and large conductance Ca2+- activated K+ channels to rat Purkinje neuron function. J. Physiol. 2003, 548, 53–69. [Google Scholar] [CrossRef] [PubMed]
  103. Gutzmann, J.J.; Lin, L.; Hoffman, D.A. Functional Coupling of Cav2.3 and BK Potassium Channels Regulates Action Potential Repolarization and Short-Term Plasticity in the Mouse Hippocampus. Front. Cell. Neurosci. 2019, 13, 27. [Google Scholar] [CrossRef] [PubMed]
  104. Vivas, O.; Moreno, C.M.; Santana, L.F.; Hille, B. Proximal clustering between BK and CaV1.3 channels promotes functional coupling and BK channel activation at low voltage. Elife 2017, 6, e28029. [Google Scholar] [CrossRef]
  105. Smith, M.R.; Nelson, A.B.; du Lac, S. Regulation of Firing Response Gain by Calcium-Dependent Mechanisms in Vestibular Nucleus Neurons. J. Neurophysiol. 2002, 87, 2031–2042. [Google Scholar] [CrossRef]
  106. Loane, D.J.; Hicks, G.A.; Perrino, B.A.; Marrion, N.V. Inhibition of BKCa channel activity by association with calcineurin in rat brain. Eur. J. Neurosci. 2006, 24, 433–441. [Google Scholar] [CrossRef]
  107. Sun, X.; Gu, X.Q.; Haddad, G.G. Calcium Influx via L- and N-Type Calcium Channels Activates a Transient Large-Conductance Ca2+-Activated K+Current in Mouse Neocortical Pyramidal Neurons. J. Neurosci. 2003, 23, 3639–3648. [Google Scholar] [CrossRef]
  108. Plante, A.E.; Whitt, J.P.; Meredith, A.L. BK channel activation by L-type Ca2+ channels CaV1.2 and CaV1.3 during the subthreshold phase of an action potential. J. Neurophysiol. 2021, 126, 427–439. [Google Scholar] [CrossRef]
  109. Davies, P.J.; Ireland, D.R.; McLachlan, E.M. Sources of Ca2+ for different Ca2+-activated K+ conductances in neurones of the rat superior cervical ganglion. J. Physiol. 1996, 495, 353–366. [Google Scholar] [CrossRef]
  110. Womack, M.D.; Chevez, C.; Khodakhah, K. Calcium-Activated Potassium Channels Are Selectively Coupled to P/Q-Type Calcium Channels in Cerebellar Purkinje Neurons. J. Neurosci. 2004, 24, 8818–8822. [Google Scholar] [CrossRef]
  111. Isaacson, J.S.; Murphy, G. Glutamate-Mediated Extrasynaptic Inhibition: Direct Coupling of NMDA Receptors to Ca2+-Activated K+ Channels. Neuron 2001, 31, 1027–1034. [Google Scholar] [CrossRef]
  112. Gómez, R.; Maglio, L.E.; Gonzalez-Hernandez, A.J.; Rivero-Pérez, B.; Bartolomé-Martín, D.; Giraldez, T. NMDA receptor–BK channel coupling regulates synaptic plasticity in the barrel cortex. Proc. Natl. Acad. Sci. USA 2021, 118, e2107026118. [Google Scholar] [CrossRef]
  113. Irie, T.; Trussell, L.O. Double-Nanodomain Coupling of Calcium Channels, Ryanodine Receptors, and BK Channels Controls the Generation of Burst Firing. Neuron 2017, 96, 856–870.e4. [Google Scholar] [CrossRef] [Green Version]
  114. Wang, B.; Bugay, V.; Ling, L.; Chuang, H.H.; Jaffe, D.B.; Brenner, R. Knockout of the BK β4-subunit promotes a functional coupling of BK channels and ryanodine receptors that mediate a fAHP-induced increase in excitability. J. Neurophysiol. 2016, 116, 456–465. [Google Scholar] [CrossRef]
  115. Whitt, J.P.; McNally, B.A.; Meredith, A.L. Differential contribution of Ca2+ sources to day and night BK current activation in the circadian clock. J. Gen. Physiol. 2017, 150, 259–275. [Google Scholar] [CrossRef]
  116. Wu, Y.; Liu, Y.; Hou, P.; Yan, Z.; Kong, W.; Liu, B.; Li, X.; Yao, J.; Zhang, Y.; Qin, F.; et al. TRPV1 Channels Are Functionally Coupled with BK(mSlo1) Channels in Rat Dorsal Root Ganglion (DRG) Neurons. PLoS ONE 2013, 8, e78203. [Google Scholar] [CrossRef]
  117. Lin, M.T.; Hessinger, D.A.; Pearce, W.J.; Longo, L.D. Modulation of BK channel calcium affinity by differential phosphorylation in developing ovine basilar artery myocytes. Am. J. Physiol. Circ. Physiol. 2006, 291, H732–H740. [Google Scholar] [CrossRef]
  118. Yan, J.; Olsen, J.V.; Park, K.-S.; Li, W.; Bildl, W.; Schulte, U.; Aldrich, R.W.; Fakler, B.; Trimmer, J.S. Profiling the Phospho-status of the BKCa Channel α Subunit in Rat Brain Reveals Unexpected Patterns and Complexity. Mol. Cell. Proteom. 2008, 7, 2188–2198. [Google Scholar] [CrossRef]
  119. Wang, Z.-W. Regulation of Synaptic Transmission by Presynaptic CaMKII and BK Channels. Mol. Neurobiol. 2008, 38, 153–166. [Google Scholar] [CrossRef]
  120. Wang, J.; Zhou, Y.; Wen, H.; Levitan, I.B. Simultaneous Binding of Two Protein Kinases to a Calcium-Dependent Potassium Channel. J. Neurosci. 1999, 19, RC4. [Google Scholar] [CrossRef]
  121. Lin, M.T.; Longo, L.D.; Pearce, W.J.; Hessinger, D.A. Ca2+-activated K+ channel-associated phosphatase and kinase activities during development. Am. J. Physiol.-Hear Circ. Physiol. 2005, 289, 414–425. [Google Scholar] [CrossRef] [PubMed]
  122. Shelley, C.; Whitt, J.P.; Montgomery, J.R.; Meredith, A.L. Phosphorylation of a constitutive serine inhibits BK channel variants containing the alternate exon “SRKR”. J. Gen. Physiol. 2013, 142, 585–598. [Google Scholar] [CrossRef] [PubMed]
  123. Jin, P.; Weiger, T.M.; Wu, Y.; Levitan, I.B. Phosphorylation-dependent Functional Coupling of hSlo Calcium-dependent Potassium Channel and Its hβ4 Subunit. J. Biol. Chem. 2002, 277, 10014–10020. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Nelson, A.B.; Gittis, A.; du Lac, S. Decreases in CaMKII Activity Trigger Persistent Potentiation of Intrinsic Excitability in Spontaneously Firing Vestibular Nucleus Neurons. Neuron 2005, 46, 623–631. [Google Scholar] [CrossRef]
  125. van Welie, I.; du Lac, S. Bidirectional control of BK channel open probability by CAMKII and PKC in medial vestibular nucleus neurons. J. Neurophysiol. 2011, 105, 1651–1659. [Google Scholar] [CrossRef]
  126. Widmer, H.A.; Rowe, I.C.M.; Shipston, M.J.; D’Antona, G.; Pellegrino, M.A.; Adami, R.; Rossi, R.; Carlizzi, C.N.; Canepari, M.; Saltin, B.; et al. Conditional protein phosphorylation regulates BK channel activity in rat cerebellar Purkinje neurons. J. Physiol. 2003, 552, 379–391. [Google Scholar] [CrossRef]
  127. Watanave, M.; Takahashi, N.; Hosoi, N.; Konno, A.; Yamamoto, H.; Yasui, H.; Kawachi, M.; Horii, T.; Matsuzaki, Y.; Hatada, I.; et al. Protein kinase Cγ in cerebellar Purkinje cells regulates Ca2+-activated large-conductance K+ channels and motor coordination. Proc. Natl. Acad. Sci. USA 2022, 119, 1–11. [Google Scholar] [CrossRef]
  128. Liu, Q.; Chen, B.; Ge, Q.; Wang, Z.W. Presynaptic Ca2+/calmodulin-dependent protein kinase II modulates neurotransmitter release by activating BK channels at Caenorhabditis elegans neuromuscular junction. J. Neurosci. 2007, 27, 10404–10413. [Google Scholar] [CrossRef]
  129. Zaman, T.; De Oliveira, C.; Smoka, M.; Narla, C.; Poulter, M.O.; Schmid, S. BK Channels Mediate Synaptic Plasticity Underlying Habituation in Rats. J. Neurosci. 2017, 37, 4540–4551. [Google Scholar] [CrossRef]
  130. Sansom, S.C.; Stockand, J.D.; Hall, D.; Williams, B. Regulation of Large Calcium-activated Potassium Channels by Protein Phosphatase 2A. J. Biol. Chem. 1997, 272, 9902–9906. [Google Scholar] [CrossRef]
  131. Tian, L.; Knaus, H.-G.; Shipston, M. Glucocorticoid Regulation of Calcium-activated Potassium Channels Mediated by Serine/Threonine Protein Phosphatase. J. Biol. Chem. 1998, 273, 13531–13536. [Google Scholar] [CrossRef]
  132. Zhou, X.-B.; Ruth, P.; Schlossmann, J.; Hofmann, F.; Korth, M. Protein Phosphatase 2A Is Essential for the Activation of Ca2+-activated K+ Currents by cGMP-dependent Protein Kinase in Tracheal Smooth Muscle and Chinese Hamster Ovary Cells. J. Biol. Chem. 1996, 271, 19760–19767. [Google Scholar] [CrossRef]
  133. Petrik, D.; Wang, B.; Brenner, R. Modulation by the BK accessory β4 subunit of phosphorylation-dependent changes in excitability of dentate gyrus granule neurons. Eur. J. Neurosci. 2011, 34, 695–704. [Google Scholar] [CrossRef] [Green Version]
  134. Ramírez-Latorre, J.A. Functional Upregulation of Ca2+ -Activated K+ Channels in the Development of Substantia Nigra Dopamine Neurons. PLoS ONE 2012, 7, e51610. [Google Scholar] [CrossRef]
  135. Kang, J.; Huguenard, J.R.; Prince, D.A. Development of BK channels in neocortical pyramidal neurons. J. Neurophysiol. 1996, 76, 188–198. [Google Scholar] [CrossRef]
  136. Kros, C.J.; Ruppersberg, J.P.; Rüsch, A. Expression of a potassium current in inner hair cells during development of hearing in mice. Nature 1998, 394, 281–284. [Google Scholar] [CrossRef]
  137. Hunsberger, M.S.; Mynlieff, M. BK potassium currents contribute differently to action potential waveform and firing rate as rat hippocampal neurons mature in the first postnatal week. J. Neurophysiol. 2020, 124, 703–714. [Google Scholar] [CrossRef]
  138. Niday, Z.; Bean, B.P. BK Channel Regulation of Afterpotentials and Burst Firing in Cerebellar Purkinje Neurons. J. Neurosci. 2021, 41, 2854–2869. [Google Scholar] [CrossRef]
  139. Kaufmann, W.; Kasugai, Y.; Ferraguti, F.; Storm, J. Two distinct pools of large-conductance calcium-activated potassium channels in the somatic plasma membrane of central principal neurons. Neuroscience 2010, 169, 974–986. [Google Scholar] [CrossRef]
  140. Womack, M.D.; Khodakhah, K. Characterization of large conductance Ca2+-activated K+ channels in cerebellar Purkinje neurons. Eur. J. Neurosci. 2002, 16, 1214–1222. [Google Scholar] [CrossRef]
  141. Sánchez-Aguilera, A.; Monedero, G.; Colino, A.; Vicente-Torres, M. Development of Action Potential Waveform in Hippocampal CA1 Pyramidal Neurons. Neuroscience 2020, 442, 151–167. [Google Scholar] [CrossRef] [PubMed]
  142. Gu, N.; Vervaeke, K.; Storm, J.F. BK potassium channels facilitate high-frequency firing and cause early spike frequency adaptation in rat CA1 hippocampal pyramidal cells. J. Physiol. 2007, 580, 859–882. [Google Scholar] [CrossRef] [PubMed]
  143. Shao, L.R.; Halvorsrud, R.; Borg-Graham, L.; Storm, J.F. The role of BK-type Ca2+-dependent K+ channels in spike broadening during repetitive firing in rat hippocampal pyramidal cells. J. Physiol. 1999, 521, 135–146. [Google Scholar] [CrossRef] [PubMed]
  144. Benhassine, N.; Berger, T. Large-conductance calcium-dependent potassium channels prevent dendritic excitability in neocortical pyramidal neurons. Pflügers Arch.-Eur. J. Physiol. 2008, 457, 1133–1145. [Google Scholar] [CrossRef]
  145. Kurowski, P.; Grzelka, K.; Szulczyk, P. Ionic Mechanism Underlying Rebound Depolarization in Medial Prefrontal Cortex Pyramidal Neurons. Front. Cell. Neurosci. 2018, 12, 93. [Google Scholar] [CrossRef]
  146. Galarraga, E.; Vilchis, C.; Tkatch, T.; Salgado, H.; Tecuapetla, F.; Perez-Rosello, T.; Perez-Garci, E.; Hernandez-Echeagaray, E.; Surmeier, D.; Bargas, J. Somatostatinergic modulation of firing pattern and calcium-activated potassium currents in medium spiny neostriatal neurons. Neuroscience 2007, 146, 537–554. [Google Scholar] [CrossRef]
  147. Vilchis, C.; Bargas, J.; Ayala, G.; Galván, E.; Galarraga, E. Ca2+ channels that activate Ca2+-dependent K+ currents in neostriatal neurons. Neuroscience 1999, 95, 745–752. [Google Scholar] [CrossRef]
  148. Wang, H.; Ardiles, A.O.; Yang, S.; Tran, T.; Posada-Duque, R.; Valdivia, G.; Baek, M.; Chuang, Y.-A.; Palacios, A.G.; Gallagher, M.; et al. Metabotropic Glutamate Receptors Induce a Form of LTP Controlled by Translation and Arc Signaling in the Hippocampus. J. Neurosci. 2016, 36, 1723–1729. [Google Scholar] [CrossRef]
  149. Kohashi, T.; Carlson, B.A. A fast BK-type KCa current acts as a postsynaptic modulator of temporal selectivity for communication signals. Front. Cell. Neurosci. 2014, 8, 286. [Google Scholar] [CrossRef]
  150. Klyachko, V.A.; Ahern, G.P.; Jackson, M.B. cGMP-mediated facilitation in nerve terminals by enhancement of the spike afterhyperpolarization. Neuron 2001, 31, 1015–1025. [Google Scholar] [CrossRef]
  151. Zhang, Y.; Bonnan, A.; Bony, G.; Ferezou, I.; Pietropaolo, S.; Ginger, M.; Sans, N.; Rossier, J.; Oostra, B.; LeMasson, G.; et al. Dendritic channelopathies contribute to neocortical and sensory hyperexcitability in Fmr1-/y mice. Nat. Neurosci. 2014, 17, 1701–1709. [Google Scholar] [CrossRef]
  152. Roshchin, M.V.; Matlashov, M.E.; Ierusalimsky, V.N.; Balaban, P.M.; Belousov, V.V.; Kemenes, G.; Staras, K.; Nikitin, E.S. A BK channel–mediated feedback pathway links single-synapse activity with action potential sharpening in repetitive firing. Sci. Adv. 2018, 4, 1–9. [Google Scholar] [CrossRef]
  153. Książek, A.; Ładno, W.; Szulczyk, B.P.; Grzelka, K.; Szulczyk, P.J. Properties of BK-type Ca++-dependent K+ channel currents in medial prefrontal cortex pyramidal neurons in rats of different ages. Front. Cell. Neurosci. 2013, 7, 185. [Google Scholar] [CrossRef] [Green Version]
  154. Benhassine, N.; Berger, T. Homogeneous distribution of large-conductance calcium-dependent potassium channels on soma and apical dendrite of rat neocortical layer 5 pyramidal neurons. Eur. J. Neurosci. 2005, 21, 914–926. [Google Scholar] [CrossRef]
  155. Dong, P.; Zhang, Y.; Hunanyan, A.S.; Mikati, M.A.; Cui, J.; Yang, H. Neuronal mechanism of a BK channelopathy in absence epilepsy and dyskinesia. Proc. Natl. Acad. Sci. USA 2022, 119, e2200140119. [Google Scholar] [CrossRef]
  156. Grewe, B.F.; Bonnan, A.; Frick, A. Back-propagation of physiological action potential output in dendrites of slender-tufted L5A pyramidal neurons. Front. Cell. Neurosci. 2010, 4, 13. [Google Scholar] [CrossRef]
  157. Sailer, C.A.; Kaufmann, W.A.; Kogler, M.; Chen, L.; Sausbier, U.; Ottersen, O.P.; Ruth, P.; Shipston, M.J.; Knaus, H.-G. Immunolocalization of BK channels in hippocampal pyramidal neurons. Eur. J. Neurosci. 2006, 24, 442–454. [Google Scholar] [CrossRef]
  158. Golding, N.L.; Jung, H.-Y.; Mickus, T.; Spruston, N. Dendritic Calcium Spike Initiation and Repolarization Are Controlled by Distinct Potassium Channel Subtypes in CA1 Pyramidal Neurons. J. Neurosci. 1999, 19, 8789–8798. [Google Scholar] [CrossRef]
  159. Shanley, L.J.; Irving, A.J.; Rae, M.G.; Ashford, M.L.; Harvey, J. Leptin inhibits rat hippocampal neurons via activation of large conductance calcium-activated K+ channels. Nat. Neurosci. 2002, 5, 299–300. [Google Scholar] [CrossRef]
  160. Springer, S.J.; Burkett, B.J.; Schrader, L.A. Modulation of BK channels contributes to activity-dependent increase of excitability through MTORC1 activity in CA1 pyramidal cells of mouse hippocampus. Front. Cell. Neurosci. 2015, 8, 451. [Google Scholar] [CrossRef]
  161. Misonou, H.; Menegola, M.; Buchwalder, L.; Park, E.W.; Meredith, A.; Rhodes, K.J.; Aldrich, R.W.; Trimmer, J.S. Immunolocalization of the Ca2+-activated K+channel Slo1 in axons and nerve terminals of mammalian brain and cultured neurons. J. Comp. Neurol. 2006, 496, 289–302. [Google Scholar] [CrossRef] [PubMed]
  162. Shruti, S.; Urban-Ciecko, J.; Fitzpatrick, J.A.; Brenner, R.; Bruchez, M.P.; Barth, A.L. The brain-specific beta4 subunit downregulates BK channel cell surface expression. PLoS ONE 2012, 7, e33429. [Google Scholar] [CrossRef]
  163. Jalini, S.; Ye, H.; Tonkikh, A.A.; Charlton, M.P.; Carlen, P.L. Raised Intracellular Calcium Contributes to Ischemia-Induced Depression of Evoked Synaptic Transmission. PLoS ONE 2016, 11, e0148110. [Google Scholar] [CrossRef] [PubMed]
  164. Alle, H.; Kubota, H.; Geiger, J.R.P. Sparse But Highly Efficient Kv3 Outpace BKCa Channels in Action Potential Repolarization at Hippocampal Mossy Fiber Boutons. J. Neurosci. 2011, 31, 8001–8012. [Google Scholar] [CrossRef]
  165. Wang, D.; Wang, X.; Liu, P.; Jing, S.; Du, H.; Zhang, L.; Jia, F.; Li, A. Serotonergic afferents from the dorsal raphe decrease the excitability of pyramidal neurons in the anterior piriform cortex. Proc. Natl. Acad. Sci. USA 2020, 117, 3239–3247. [Google Scholar] [CrossRef]
  166. Casale, A.E.; Foust, A.J.; Bal, T.; McCormick, D.A. Cortical Interneuron Subtypes Vary in Their Axonal Action Potential Properties. J. Neurosci. 2015, 35, 15555–15567. [Google Scholar] [CrossRef]
  167. Ji, J.J.; Chen, L.; Duan, X.; Song, X.; Su, W.; Zhang, P.; Li, L.; Bai, S.; Sun, Y.; Inagaki, N. BK Channels Reveal Novel Phosphate Sensitivity in SNr Neurons. PLoS ONE 2012, 7, e52148. [Google Scholar] [CrossRef]
  168. Authement, M.E.; Langlois, L.D.; Shepard, R.D.; Browne, C.A.; Lucki, I.; Kassis, H.; Nugent, F.S. A role for corticotropin-releasing factor signaling in the lateral habenula and its modulation by early-life stress. Sci. Signal. 2018, 11, eaan6480. [Google Scholar] [CrossRef]
  169. Zhang, Y.Y.; Han, X.; Liu, Y.; Chen, J.; Hua, L.; Ma, Q.; Huang, Y.Y.X.; Tang, Q.Y.; Zhang, Z. +MRNA expression of LRRC55 protein (leucine-rich repeat-containing protein 55) in the adult mouse brain. PLoS ONE 2018, 13, e0191749. [Google Scholar]
  170. Ster, J.; De Bock, F.; Guérineau, N.C.; Janossy, A.; Barrère-Lemaire, S.; Bos, J.L.; Bockaert, J.; Fagni, L. Exchange protein activated by cAMP (Epac) mediates cAMP activation of p38 MAPK and modulation of Ca2+-dependent K+ channels in cerebellar neurons. Proc. Natl. Acad. Sci. USA 2007, 104, 2519–2524. [Google Scholar] [CrossRef]
  171. Rancz, E.A.; Häusser, M. Dendritic calcium spikes are tunable triggers of cannabinoid release and short-term synaptic plasticity in cerebellar purkinje neurons. J. Neurosci. 2006, 26, 5428–5437. [Google Scholar] [CrossRef]
  172. Ouares, K.A.; Filipis, L.; Tzilivaki, A.; Poirazi, P.; Canepari, M. Two Distinct Sets of Ca2+ and K+ Channels Are Activated at Different Membrane Potentials by the Climbing Fiber Synaptic Potential in Purkinje Neuron Dendrites. J. Neurosci. 2019, 39, 1969–1981. [Google Scholar] [CrossRef]
  173. Hirono, M.; Ogawa, Y.; Misono, K.; Zollinger, D.R.; Trimmer, J.S.; Rasband, M.N.; Misonou, H. BK Channels Localize to the Paranodal Junction and Regulate Action Potentials in Myelinated Axons of Cerebellar Purkinje Cells. J. Neurosci. 2015, 35, 7082–7094. [Google Scholar] [CrossRef]
  174. Montgomery, J.R.; Meredith, A.L. Genetic activation of BK currents in vivo generates bidirectional effects on neuronal excitability. Proc. Natl. Acad. Sci. USA 2012, 109, 18997–19002. [Google Scholar] [CrossRef] [Green Version]
  175. He, S.; Wang, Y.-X.; Petralia, R.S.; Brenowitz, S.D. Cholinergic Modulation of Large-Conductance Calcium-Activated Potassium Channels Regulates Synaptic Strength and Spine Calcium in Cartwheel Cells of the Dorsal Cochlear Nucleus. J. Neurosci. 2014, 34, 5261–5272. [Google Scholar] [CrossRef]
  176. Garaycochea, J.; Slaughter, M.M. GABAB receptors enhance excitatory responses in isolated rat retinal ganglion cells. J. Physiol. 2016, 594, 5543–5554. [Google Scholar] [CrossRef]
  177. Womack, M.; Hoang, C.; Khodakhah, K. Large conductance calcium-activated potassium channels affect both spontaneous firing and intracellular calcium concentration in cerebellar Purkinje neurons. Neuroscience 2009, 162, 989–1000. [Google Scholar] [CrossRef]
  178. Nelson, A.B.; Krispel, C.M.; Sekirnjak, C.; du Lac, S. Long-Lasting Increases in Intrinsic Excitability Triggered by Inhibition. Neuron 2003, 40, 609–620. [Google Scholar] [CrossRef]
  179. Kimm, T.; Khaliq, Z.M.; Bean, B.P. Differential Regulation of Action Potential Shape and Burst-Frequency Firing by BK and Kv2 Channels in Substantia Nigra Dopaminergic Neurons. J. Neurosci. 2015, 35, 16404–16417. [Google Scholar] [CrossRef]
  180. Gittis, A.; Moghadam, S.H.; Du Lac, S. Mechanisms of Sustained High Firing Rates in Two Classes of Vestibular Nucleus Neurons: Differential Contributions of Resurgent Na, Kv3, and BK Currents. J. Neurophysiol. 2010, 104, 1625–1634. [Google Scholar] [CrossRef]
  181. Petrucci, C.; Resta, V.; Fieni, F.; Bigiani, A.; Bagnoli, P. Modulation of potassium current and calcium influx by somatostatin in rod bipolar cells isolated from the rabbit retina via sst2 receptors. Naunyn-Schmiedeberg’s Arch. Pharmacol. 2001, 363, 680–694. [Google Scholar] [CrossRef] [PubMed]
  182. Gittis, A.H.; Lac, S. Firing Properties of GABAergic Versus Non-GABAergic Vestibular Nucleus Neurons Conferred by a Differential Balance of Potassium Currents. J. Neurophysiol. 2007, 2003, 3986. [Google Scholar] [CrossRef] [PubMed]
  183. Kohlmeier, K.A.; Christensen, M.H.; Kristensen, M.P.; Kristiansen, U. Pharmacological evidence of functional inhibitory metabotrophic glutamate receptors on mouse arousal-related cholinergic laterodorsal tegmental neurons. Neuropharmacology 2013, 66, 99–113. [Google Scholar] [CrossRef] [PubMed]
  184. Liu, Y.; Savtchouk, I.; Acharjee, S.; Liu, S.J. Inhibition of Ca2+-activated large-conductance K+ channel activity alters synaptic AMPA receptor phenotype in mouse cerebellar stellate cells. J. Neurophysiol. 2011, 106, 144–152. [Google Scholar] [CrossRef] [Green Version]
  185. Bock, T.; Stuart, G.J. Impact of calcium-activated potassium channels on NMDA spikes in cortical layer 5 pyramidal neurons. J. Neurophysiol. 2016, 115, 1740–1748. [Google Scholar] [CrossRef]
  186. Golding, N.L.; Staff, N.P.; Spruston, N. Dendritic spikes as a mechanism for cooperative long-term potentiation. Nature 2002, 418, 326–331. [Google Scholar] [CrossRef]
  187. Poolos, N.P.; Johnston, D. Calcium-Activated Potassium Conductances Contribute to Action Potential Repolarization at the Soma But Not the Dendrites of Hippocampal CA1 Pyramidal Neurons. J. Neurosci. 1999, 19, 5205–5212. [Google Scholar] [CrossRef]
  188. Egaña-Huguet, J.; Saumell-Esnaola, M.; Achicallende, S.; Soria-Gomez, E.; Río, I.B.-D.; del Caño, G.G.; Barrondo, S.; Sallés, J.; Gerrikagoitia, I.; Puente, N.; et al. Lack of the Transient Receptor Potential Vanilloid 1 Shifts Cannabinoid-Dependent Excitatory Synaptic Plasticity in the Dentate Gyrus of the Mouse Brain Hippocampus. Front. Neuroanat. 2021, 15, 701573. [Google Scholar] [CrossRef]
  189. Sclip, A.; Acuna, C.; Luo, F.; Südhof, T.C. RIM-binding proteins recruit BK-channels to presynaptic release sites adjacent to voltage-gated Ca2+ -channels. EMBO J. 2018, 37, 1–14. [Google Scholar] [CrossRef]
  190. Brenes, O.; Vandael, D.; Carbone, E.; Montarolo, P.; Ghirardi, M. Knock-down of synapsin alters cell excitability and action potential waveform by potentiating BK and voltage-gated Ca2+ currents in Helix serotonergic neurons. Neuroscience 2015, 311, 430–443. [Google Scholar] [CrossRef]
  191. Oh, K.H.; Abraham, L.S.; Gegg, C.; Silvestri, C.; Huang, Y.-C.; Alkema, M.J.; Furst, J.; Raicu, D.; Kim, H. Presynaptic BK channel localization is dependent on the hierarchical organization of alpha-catulin and dystrobrevin and fine-tuned by CaV2 calcium channels. BMC Neurosci. 2015, 16, 1–13. [Google Scholar] [CrossRef]
  192. Chen, B.; Liu, P.; Zhan, H.; Wang, Z.-W. Dystrobrevin Controls Neurotransmitter Release and Muscle Ca2+ Transients by Localizing BK Channels in Caenorhabditis elegans. J. Neurosci. 2011, 31, 17338–17347. [Google Scholar] [CrossRef]
  193. Abrahamsson, T.; Chou, C.Y.C.; Li, S.Y.; Mancino, A.; Costa, R.P.; Brock, J.A.; Nuro, E.; Buchanan, K.A.; Elgar, D.; Blackman, A.V.; et al. Differential Regulation of Evoked and Spontaneous Release by Presynaptic NMDA Receptors. Neuron 2017, 96, 839–855.e5. [Google Scholar] [CrossRef]
  194. Banerjee, A.; Larsen, R.; Philpot, B.D.; Paulsen, O. Roles of Presynaptic NMDA Receptors in Neurotransmission and Plasticity. Trends Neurosci. 2015, 39, 26–39. [Google Scholar] [CrossRef] [Green Version]
  195. Lituma, P.J.; Kwon, H.-B.; Alviña, K.; Luján, R.; Castillo, P.E. Presynaptic NMDA receptors facilitate short-term plasticity and BDNF release at hippocampal mossy fiber synapses. Elife 2021, 10, e66612. [Google Scholar] [CrossRef]
  196. Martire, M.; Barrese, V.; D’Amico, M.; Iannotti, F.A.; Pizzarelli, R.; Samengo, I.; Viggiano, D.; Ruth, P.; Cherubini, E.; Taglialatela, M. Pre-synaptic BK channels selectively control glutamate versus GABA release from cortical and hippocampal nerve terminals. J. Neurochem. 2010, 115, 411–422. [Google Scholar] [CrossRef]
  197. Samengo, I.; Curro’, D.; Barrese, V.; Taglialatela, M.; Martire, M. Large Conductance Calcium-Activated Potassium Channels: Their Expression and Modulation of Glutamate Release from Nerve Terminals Isolated from Rat Trigeminal Caudal Nucleus and Cerebral Cortex. Neurochem. Res. 2014, 39, 901–910. [Google Scholar] [CrossRef]
  198. Shoudai, K.; Nonaka, K.; Maeda, M.; Wang, Z.-M.; Jeong, H.-J.; Higashi, H.; Murayama, N.; Akaike, N. Effects of various K+ channel blockers on spontaneous glycine release at rat spinal neurons. Brain Res. 2007, 1157, 11–22. [Google Scholar] [CrossRef]
  199. De San Martín, J.Z.; Pyott, S.; Ballestero, J.; Katz, E. Ca2+ and Ca2+-activated K+ channels that support and modulate transmitter release at the olivocochlear efferent-inner hair cell synapse. J. Neurosci. 2010, 30, 12157–12167. [Google Scholar] [CrossRef]
  200. Choi, T.-Y.; Lee, S.-H.; Kim, Y.-J.; Bae, J.R.; Lee, K.M.; Jo, Y.; Kim, S.-J.; Lee, A.-R.; Choi, S.; Choi, L.-M.; et al. Cereblon Maintains Synaptic and Cognitive Function by Regulating BK Channel. J. Neurosci. 2018, 38, 3571–3583. [Google Scholar] [CrossRef]
  201. Price, D.L.; Ludwig, J.W.; Mi, H.; Schwarz, T.L.; Ellisman, M.H. Distribution of rSlo Ca2+-activated K+ channels in rat astrocyte perivascular endfeet. Brain Res. 2002, 956, 183–193. [Google Scholar] [CrossRef] [PubMed]
  202. Girouard, H.; Bonev, A.D.; Hannah, R.M.; Meredith, A.; Aldrich, R.W.; Nelson, M.T. Astrocytic endfoot Ca 2+ and BK channels determine both arteriolar dilation and constriction. Proc. Natl. Acad. Sci. USA 2010, 107, 3811–3816. [Google Scholar] [CrossRef] [PubMed]
  203. Kyle, B.; Mishra, R.C.; Braun, A.P. The augmentation of BK channel activity by nitric oxide signaling in rat cerebral arteries involves co-localized regulatory elements. J. Cereb. Blood Flow Metab. 2017, 37, 3759–3773. [Google Scholar] [CrossRef] [PubMed]
  204. Filosa, J.A.; Bonev, A.D.; Straub, S.V.; Meredith, A.; Wilkerson, M.K.; Aldrich, R.W.; Nelson, M. Local potassium signaling couples neuronal activity to vasodilation in the brain. Nat. Neurosci. 2006, 9, 1397–1403. [Google Scholar] [CrossRef]
  205. Gebremedhin, D.; Yamaura, K.; Zhang, C.; Bylund, J.; Koehler, R.C.; Harder, D.R. Metabotropic Glutamate Receptor Activation Enhances the Activities of Two Types of Ca2+-Activated K+Channels in Rat Hippocampal Astrocytes. J. Neurosci. 2003, 23, 1678–1687. [Google Scholar] [CrossRef]
  206. Brandalise, F.; Ratto, D.; Leone, R.; Olivero, F.; Roda, E.; Locatelli, C.A.; Grazia Bottone, M.; Rossi, P. Deeper and Deeper on the Role of BK and Kir4.1 Channels in Glioblastoma Invasiveness: A Novel Summative Mechanism? Front Neurosci. 2020, 14, 1–6. [Google Scholar] [CrossRef]
  207. Singh, H.; Li, M.; Hall, L.; Chen, S.; Sukur, S.; Lu, R.; Caputo, A.; Meredith, A.; Stefani, E.; Toro, L. MaxiK channel interactome reveals its interaction with GABA transporter 3 and heat shock protein 60 in the mammalian brain. Neuroscience 2016, 317, 76–107. [Google Scholar] [CrossRef]
  208. Hayashi, Y.; Nakanishi, H. BK channel in microglia as a potent therapeutic molecular target for neuropathic pain. J. Oral Biosci. 2015, 57, 131–134. [Google Scholar] [CrossRef]
  209. Huang, S.; Chen, T.; Suo, Q.; Shi, R.; Khan, H.; Ma, Y.; Tang, Y.; Yang, G.-Y.; Zhang, Z. BK Channel-Mediated Microglial Phagocytosis Alleviates Neurological Deficit After Ischemic Stroke. Front. Cell. Neurosci. 2021, 15, 683769. [Google Scholar] [CrossRef]
  210. Hayashi, Y.; Kawaji, K.; Sun, L.; Zhang, X.; Koyano, K.; Yokoyama, T.; Kohsaka, S.; Inoue, K.; Nakanishi, H. Microglial Ca2+-Activated K+ Channels Are Possible Molecular Targets for the Analgesic Effects of S-Ketamine on Neuropathic Pain. J. Neurosci. 2011, 31, 17370–17382. [Google Scholar] [CrossRef]
  211. Bordey, A.; Spencer, D.D. Chemokine modulation of high-conductance Ca2+-sensitive K+ currents in microglia from human hippocampi. J. FENS 2003, 18, 2893–2898. [Google Scholar] [CrossRef]
  212. Schilling, T.; Eder, C. Ion channel expression in resting and activated microglia of hippocampal slices from juvenile mice. Brain Res. 2007, 1186, 21–28. [Google Scholar] [CrossRef]
  213. Schilling, T.; Eder, C. Microglial K + channel expression in young adult and aged mice. Glia 2014, 63, 664–672. [Google Scholar] [CrossRef]
  214. Buttigieg, J.; Eftekharpour, E.; Karimi-Abdolrezaee, S.; Fehlings, M.G. Molecular and electrophysiological evidence for the expression of BK channels in oligodendroglial precursor cells. Eur. J. Neurosci. 2011, 34, 538–547. [Google Scholar] [CrossRef]
  215. Larson, V.A.; Zhang, Y.; Bergles, D.E. Electrophysiological properties of NG2 þ cells: Matching physiological studies with gene expression profiles. Brain Res. 2015, 15, 1853. [Google Scholar]
  216. Jäkel, S.; Agirre, E.; Falcão, A.M.; van Bruggen, D.; Lee, K.W.; Knuesel, I.; Malhotra, D.; Ffrench-Constant, C.; Williams, A.; Castelo-Branco, G. Altered human oligodendrocyte heterogeneity in multiple sclerosis. Nature 2019, 566, 543–547. [Google Scholar] [CrossRef]
  217. Valverde, M.A.; Rojas, P.; Amigo, J.; Cosmelli, D.; Orio, P.; Bahamonde, M.I.; Mann, G.E.; Vergara, C.; Latorre, R. Acute Activation of Maxi-K Channels (hSlo) by Estradiol Binding to the β Subunit. Science 1999, 285, 1929–1931. [Google Scholar] [CrossRef]
  218. King, J.T.; Lovell, P.V.; Rishniw, M.; Kotlikoff, M.I.; Zeeman, M.L.; McCobb, D.P. β2 and β4 Subunits of BK Channels Confer Differential Sensitivity to Acute Modulation by Steroid Hormones. J. Neurophysiol. 2006, 95, 2878–2888. [Google Scholar] [CrossRef]
  219. Teyler, T.J.; Vardaris, R.M.; Lewis, D.; Rawitch, A.B. Gonadal Steroids: Effects on Excitability of Hippocampal Pyramidal Cells. Science 1980, 209, 1017–1018. [Google Scholar] [CrossRef]
  220. Oberlander, J.G.; Woolley, C.S. 17β-Estradiol Acutely Potentiates Glutamatergic Synaptic Transmission in the Hippocampus through Distinct Mechanisms in Males and Females. J. Neurosci. 2016, 36, 2677–2690. [Google Scholar] [CrossRef]
  221. Soutar, C.N.; Grenier, P.; Patel, A.; Kabitsis, P.P.; Olmstead, M.C.; Bailey, C.D.C.; Dringenberg, H.C. Brain-Generated 17β-Estradiol Modulates Long-Term Synaptic Plasticity in the Primary Auditory Cortex of Adult Male Rats. Cereb. Cortex 2021, 32, 2140–2155. [Google Scholar] [CrossRef]
  222. Hoshi, T.; Tian, Y.; Xu, R.; Heinemann, S.H.; Hou, S. Mechanism of the modulation of BK potassium channel complexes with different auxiliary subunit compositions by the omega-3 fatty acid DHA. Proc. Natl. Acad. Sci. USA 2013, 110, 4822–4827. [Google Scholar] [CrossRef] [PubMed]
  223. Sun, X.; Zhou, D.; Zhang, P.; Moczydlowski, E.G.; Haddad, G.G. β-subunit-dependent modulation of hSlo BK current by arachidonic acid. J. Neurophysiol. 2007, 97, 62–69. [Google Scholar] [CrossRef] [PubMed]
  224. Duerson, K.; White, R.E.; Jiang, F.; Schonbrunn, A.; Armstrong, D.L. Somatostatin stimulates BK(Ca) channels in rat pituitary tumor cells through lipoxygenase metabolites of arachidonic acid. Neuropharmacology 1996, 35, 949–961. [Google Scholar] [CrossRef] [PubMed]
  225. Barlow, R.S.; El-Mowafy, A.M.; White, A.R.E. H2O2 opens BK(Ca) channels via the PLA2-arachidomic acid signaling cascade in coronary artery smooth muscle. Am. J. Physiol.-Hear. Circ. Physiol. 2000, 279, H475–H483. [Google Scholar] [CrossRef] [PubMed]
  226. Bondarenko, A.I.; Panasiuk, O.; Okhai, I.; Montecucco, F.; Brandt, K.J.; Mach, F. Direct activation of Ca2+ and voltage-gated potassium channels of large conductance by anandamide in endothelial cells does not support the presence of endothelial atypical cannabinoid receptor. Eur. J. Pharmacol. 2017, 805, 14–24. [Google Scholar] [CrossRef]
  227. Sade, H.; Muraki, K.; Ohya, S.; Hatano, N.; Imaizumi, Y. Activation of large-conductance, Ca2+-activated K+ channels by cannabinoids. Am. J. Physiol.-Cell Physiol. 2006, 290, C77–C86. [Google Scholar] [CrossRef]
  228. Li, Y.; Zhang, L.; Wu, Y.; Zheng, Q.; Chen, M.; Qian, Z.; Wei, C.; Han, J.; Liu, Z.; Ren, W.; et al. Cannabinoids-induced peripheral analgesia depends on activation of BK channels. Brain Res. 2019, 1711, 23–28. [Google Scholar] [CrossRef]
  229. Bailey, C.S.; Moldenhauer, H.J.; Park, S.M.; Keros, S.; Meredith, A.L. KCNMA1-linked channelopathy. J. Gen. Physiol. 2019, 151, 1173–1189. [Google Scholar] [CrossRef]
  230. Miller, J.P.; Moldenhauer, H.J.; Keros, S.; Meredith, A.L. An emerging spectrum of variants and clinical features in KCNMA1-linked channelopathy. Channels 2021, 15, 447–464. [Google Scholar] [CrossRef]
  231. Du, W.; Bautista, J.F.; Yang, H.; Diez-Sampedro, A.; You, S.-A.; Wang, L.; Kotagal, P.; Lüders, H.O.; Shi, J.; Cui, J.; et al. Calcium-sensitive potassium channelopathy in human epilepsy and paroxysmal movement disorder. Nat. Genet. 2005, 37, 733–738. [Google Scholar] [CrossRef]
  232. Mi Park, S.; Roache, C.E.; Iffland, P.H., II; Moldenhauer, H.J.; Matychak, K.K.; Plante, A.E. BK channel properties correlate with neurobehavioral severity in three KCNMA1-linked channelopathy mouse models. Elife 2022, 11, e77953. [Google Scholar] [CrossRef]
  233. Li, X.; Poschmann, S.; Chen, Q.; Fazeli, W.; Oundjian, N.J.; Snoeijen-Schouwenaars, F.M.; Fricke, O.; Kamsteeg, E.-J.; Willemsen, M.; Wang, Q.K. De novo BK channel variant causes epilepsy by affecting voltage gating but not Ca2+ sensitivity. Eur. J. Hum. Genet. 2018, 26, 220–229. [Google Scholar] [CrossRef]
  234. Whitmire, L.E.; Ling, L.; Bugay, V.; Carver, C.M.; Timilsina, S.; Chuang, H.-H.; Jaffe, D.B.; Shapiro, M.S.; Cavazos, J.; Brenner, R. Downregulation of KCNMB4 expression and changes in BK channel subtype in hippocampal granule neurons following seizure activity. PLoS ONE 2017, 12, e0188064. [Google Scholar] [CrossRef]
  235. Lorenz, S.; Heils, A.; Kasper, J.M.; Sander, T. Allelic association of a truncation mutation of theKCNMB3 gene with idiopathic generalized epilepsy. Am. J. Med Genet. Part B Neuropsychiatr. Genet. 2007, 144B, 10–13. [Google Scholar] [CrossRef]
  236. Chen, S.R.; Cai, Y.Q.; Pan, H.L. Plasticity and emerging role of BKCa channels in nociceptive control in neuropathic pain. J Neurochem. 2009, 110, 352–362. [Google Scholar] [CrossRef]
  237. Furukawa, N.; Takasusuki, T.; Fukushima, T.; Hori, Y. Presynaptic large-conductance calcium-activated potassium channels control synaptic transmission in the superficial dorsal horn of the mouse. Neurosci. Lett. 2008, 444, 79–82. [Google Scholar] [CrossRef]
  238. Sarantopoulos, C.D.; McCallum, J.B.; Rigaud, M.; Fuchs, A.; Kwok, W.-M.; Hogan, Q.H. Opposing effects of spinal nerve ligation on calcium-activated potassium currents in axotomized and adjacent mammalian primary afferent neurons. Brain Res. 2007, 1132, 84–99. [Google Scholar] [CrossRef]
  239. Zhang, F.-X.; Gadotti, V.M.; Souza, I.A.; Chen, L.; Zamponi, G.W. BK Potassium Channels Suppress Cavα2δ Subunit Function to Reduce Inflammatory and Neuropathic Pain. Cell Rep. 2018, 22, 1956–1964. [Google Scholar] [CrossRef]
  240. Imari, K.; Harada, Y.; Zhang, J.; Mori, Y.; Hayashi, Y. KCNMB3 in spinal microglia contributes to the generation and maintenance of neuropathic pain in mice. Int. J. Mol. Med. 2019, 44, 1585–1593. [Google Scholar] [CrossRef]
  241. Typlt, M.; Mirkowski, M.; Azzopardi, E.; Ruettiger, L.; Ruth, P.; Schmid, S. Mice with Deficient BK Channel Function Show Impaired Prepulse Inhibition and Spatial Learning, but Normal Working and Spatial Reference Memory. PLoS ONE 2013, 8, e81270. [Google Scholar] [CrossRef] [PubMed]
  242. Du, X.; Carvalho-De-Souza, J.L.; Wei, C.; Carrasquel-Ursulaez, W.; Lorenzo, Y.; Gonzalez, N.; Kubota, T.; Staisch, J.; Hain, T.; Petrossian, N.; et al. Loss-of-function BK channel mutation causes impaired mitochondria and progressive cerebellar ataxia. Proc. Natl. Acad. Sci. USA 2020, 117, 6023–6034. [Google Scholar] [CrossRef] [PubMed]
  243. Kratschmer, P.; Lowe, S.A.; Buhl, E.; Chen, K.F.; Kullmann, D.M.; Pittman, A.; Hodge, J.J.L.; Jepson, J.E.C. Impaired Pre-Motor Circuit Activity and Movement in a Drosophila Model of KCNMA1-Linked Dyskinesia. Mov. Disord. 2021, 36, 1158–1169. [Google Scholar] [CrossRef] [PubMed]
  244. Gao, L.; Zhao, J.; Ardiel, E.L.; Hall, Q.; Nurrish, S.; Kaplan, J.M. Shank promotes action potential repolarization by recruiting BK channels to calcium microdomains. Elife 2022, 11, e75140. [Google Scholar] [CrossRef] [PubMed]
  245. Deng, P.-Y.; Klyachko, V.A. Channelopathies in fragile X syndrome. Nat. Rev. Neurosci. 2021, 22, 275–289. [Google Scholar] [CrossRef]
  246. Hébert, B.; Pietropaolo, S.; Même, S.; Laudier, B.; Laugeray, A.; Doisne, N.; Quartier, A.; Lefeuvre, S.; Got, L.; Cahard, D.; et al. Rescue of fragile X syndrome phenotypes in Fmr1KO mice by a BKCa channel opener molecule. Orphanet. J. Rare Dis. 2014, 9, 124. [Google Scholar] [CrossRef]
  247. Edwards, T.; Rickard, N. Pharmaco-behavioural evidence indicating a complex role for ryanodine receptor calcium release channels in memory processing for a passive avoidance task. Neurobiol. Learn. Mem. 2006, 86, 1–8. [Google Scholar] [CrossRef]
  248. Matthews, E.A.; Weible, A.P.; Shah, S.; Disterhoft, J.F. The BK-mediated fAHP is modulated by learning a hippocampus-dependent task. Proc. Natl. Acad. Sci. USA 2008, 105, 15154–15159. [Google Scholar] [CrossRef]
  249. Matthews, E.A.; Disterhoft, J.F. Blocking the BK channel impedes acquisition of trace eyeblink conditioning. Learn. Mem. 2009, 16, 106–109. [Google Scholar] [CrossRef]
  250. Ryan, D.P.; Henzel, K.S.; Pearson, B.L.; Siwek, M.E.; Papazoglou, A.; Guo, L.; Paesler, K.; Yu, M.; Müller, R.; Xie, K.; et al. A paternal methyl donor-rich diet altered cognitive and neural functions in offspring mice. Mol. Psychiatry 2017, 23, 1345–1355. [Google Scholar] [CrossRef]
  251. Choi, T.-Y.; Lee, S.-H.; Kim, S.-J.; Jo, Y.; Park, C.-S.; Choi, S.-Y. BK channel blocker paxilline attenuates thalidomide-caused synaptic and cognitive dysfunctions in mice. Sci. Rep. 2018, 8, 17653. [Google Scholar] [CrossRef]
  252. Beecham, G.W.; Schnetz-Boutaud, N.; Haines, J.L.; Pericak-Vance, M.A. CALHM1Polymorphism is not Associated with Late-onset Alzheimer Disease. Ann. Hum. Genet. 2009, 73, 379–381. [Google Scholar] [CrossRef]
  253. Beecham, G.W.; Hamilton, K.; Naj, A.C.; Martin, E.R.; Huentelman, M.; Myers, A.J.; Corneveaux, J.J.; Hardy, J.; Vonsattel, J.-P.; Younkin, S.G.; et al. Genome-Wide Association Meta-analysis of Neuropathologic Features of Alzheimer’s Disease and Related Dementias. PLOS Genet. 2014, 10, e1004606. [Google Scholar] [CrossRef]
  254. Ye, H.; Jalini, S.; Mylvaganam, S.; Carlen, P. Activation of large-conductance Ca2+-activated K+ channels depresses basal synaptic transmission in the hippocampal CA1 area in APP (swe/ind) TgCRND8 mice. Neurobiol. Aging 2010, 31, 591–604. [Google Scholar] [CrossRef]
  255. Yamamoto, K.; Ueta, Y.; Wang, L.; Yamamoto, R.; Inoue, N.; Inokuchi, K.; Aiba, A.; Yonekura, H.; Kato, N. Suppression of a neocortical potassium channel activity by intracellular amyloid-β and its rescue with homer1a. J. Neurosci. 2011, 31, 11100–11109. [Google Scholar] [CrossRef] [Green Version]
  256. Gavello, D.; Calorio, C.; Franchino, C.; Cesano, F.; Carabelli, V.; Carbone, E.; Marcantoni, A. Early Alterations of Hippocampal Neuronal Firing Induced by Abeta42. Cereb. Cortex 2018, 28, 433–446. [Google Scholar] [CrossRef]
  257. Wang, L.; Kang, H.; Li, Y.; Shui, Y.; Yamamoto, R.; Sugai, T.; Kato, N. Cognitive recovery by chronic activation of the large-conductance calcium-activated potassium channel in a mouse model of Alzheimer’s disease. Neuropharmacology 2015, 92, 8–15. [Google Scholar] [CrossRef]
Figure 1. BK Channel allosteric gating mechanism. (A) H–A allosteric model [43]. Voltage sensor (R–A), Ca2+ binding, and C–O transitions are defined by the equilibrium constants J, K, and L, respectively. Voltage sensors and Ca2+ sensors are coupled to the pore by allosteric factors D and C, respectively, and coupling between sensors is performed by E. (B) Simulated data of the Log10 of the probability of opening vs. voltage obtained using the H–A model. Notice that when all the voltage sensors are at rest the parameters L, zL, and C can be obtained.
Figure 1. BK Channel allosteric gating mechanism. (A) H–A allosteric model [43]. Voltage sensor (R–A), Ca2+ binding, and C–O transitions are defined by the equilibrium constants J, K, and L, respectively. Voltage sensors and Ca2+ sensors are coupled to the pore by allosteric factors D and C, respectively, and coupling between sensors is performed by E. (B) Simulated data of the Log10 of the probability of opening vs. voltage obtained using the H–A model. Notice that when all the voltage sensors are at rest the parameters L, zL, and C can be obtained.
Ijms 24 03407 g001
Figure 2. Structural characteristics of the BK channel. (A) The α subunit is formed by a transmembrane domain (TMD) composed of seven α helices (S0–S7) and a large carboxy terminal (CTD) containing two regulatory potassium conductance domains. (B) The BK channel is a tetramer formed by 4 α subunits in which the CTD acquires a swapped conformation and the 4 CTDs form the gating ring. (C) Top view of the BK channel. The voltage sensor domain (VSD) and the pore domain (PD) have been colored to highlight the non-swapped configuration of the VSD and the PD.
Figure 2. Structural characteristics of the BK channel. (A) The α subunit is formed by a transmembrane domain (TMD) composed of seven α helices (S0–S7) and a large carboxy terminal (CTD) containing two regulatory potassium conductance domains. (B) The BK channel is a tetramer formed by 4 α subunits in which the CTD acquires a swapped conformation and the 4 CTDs form the gating ring. (C) Top view of the BK channel. The voltage sensor domain (VSD) and the pore domain (PD) have been colored to highlight the non-swapped configuration of the VSD and the PD.
Ijms 24 03407 g002
Table 1. Functional expression of BK channels in the central nervous system.
Table 1. Functional expression of BK channels in the central nervous system.
CNS Area and Neuronal TypeSite of Neuronal ExpressionFunctionSubunit and/or PharmacologyReferences
Prefrontal cortex pyramidal neuronsPre- (AIS and axon) and postsynaptic (soma)Limit AP broadening and allow high-frequency firing.IBTX-sensitive[151,152]
Pax-sensitive[153]
Visual cortex pyramidal neuronsPostsynaptic (basal dendrites)Reduce EPSPs in small-headed spines.α subunit, Chtx-sensitive[25]
Somatosensorial layer 5 pyramidal neuronsPostsynaptic (soma and dendrites)Reduce influx of ions through NMDAR and increase threshold for plasticity.
Modulate dendritic Na+ and Ca2+ spikes.
IBTX and Pax-sensitive[154]
Pax-sensitive[112,155,156]
Hippocampal CA1 pyramidal neuronPre- (axon) and Postsynaptic (soma and apical dendrites)Limit repetitive firing of dendritic calcium spikes. AP repolarization and fAHP.α subunit[141,157]
Chtx-sensitive[158,159]
Pax and IBTX-sensitive[160]
IBTX-sensitive[137,142]
Hippocampal CA3 pyramidal neuronPre- (Axonal terminals) and postsynaptic (soma)Decrease glutamate release. Regulate neuronal excitability.α subunit[161]
β4 subunit[15,162]
Chtx-sensitive[163]
IBTX-sensitive[23]
IBTX and Pax-sensitive[24]
Hippocampal dentate gyrus granule cellsPre- (mossy fiber boutons and perforant path terminals) and postsynaptic (soma and dendrites)Control perforant path induce EPSP.
Regulate excitability.
At mossy fiber boutons contribute to AP repolarization when Kv channels are blocked.
α subunit[161]
β2 and β4 subunits[164]
β4 subunit[72]
Pax-, Chtx- and IBTX-sensitive[99]
Pax-sensitive[27]
Piriform cortex pyramidal neuronPostsynaptic (soma)Reduce excitability modulating odor perception.Pax-sensitive[165]
Anterior cingulate cortex pyramidal neuronsPre- (axonal terminals) and postsynaptic (soma)Reduce hyperexcitability induced by neuropathic pain state.β4 subunit.[73]
Somatosensorial SOM+ interneuronsPre- (axon) and postsynaptic (soma)Contribute to AP repolarization.IBTX-sensitive[166]
Amygdala central nucleus gabaergic neuronsPresynapticLimit GABA release.β2 and β4 subunits[12]
Substantia nigra pars reticulata gabaergic neuronsPostsynaptic (soma)Contribute to hyperpolarization.Chtx- and Pax-sensitive[167]
Substantia nigra dopaminergic neuronsPostsynaptic (soma)Not tested.IBTX-senstive[134]
Infragranular pyramidal neuronsPostsynaptic (soma and apical dendrites)Not tested.Chtx-sensitive[135]
Lateral habenula nucleus neuronsPostsynaptic (soma)fAHP increased by corticotropin-releasing factor activation.IBTX-sensitive[168]
Medial habenula nucleus neuronsNot describedNot tested.LRRC55 γ subunit[169]
Cerebellum Purkinje neuronsPresynaptic (paranodal junctions of axons)
Postsynaptic (soma and dendrites)
Decrease dendritic Ca2+ spikes avoiding DSE propagation.
Regulate afterdepolarization.
In myelinated axons allow high-fidelity firing of AP.
α subunit[161]
β4 subunit[170]
Penitrem A sensitive.[171]
IBTX-sensitive[172]
IBTX–sensitive + IBTX-insensitive component[138]
IBTX and Pax-sensitive.[102]
Pax-sensitive[173]
Cerebellar basket cellsPresynaptic (axonal terminals)Not tested.α subunit[161]
Cerebellar granule cellsPostsynaptic (soma)Contribute to hyperpolarization and AHP.IBTX-sensitive[170]
Suprachiasmatic nucleus neuronsPostsynaptic (soma)Increase neuronal firing at night but decrease firing during day.β4 subunit[174]
β2 subunit[86]
Supraoptic nucleus neuronsPre- and postsynaptic (dendrites, soma and nerve terminals)β4 but not β1 allows potentiation by ethanol.β1 subunit in soma and dendrites; β4 in nerve terminals[79]
Medial vestibular nucleus neuronsPostsynaptic
(soma)
Contribute to excitability. Can be activated by CaMKII and inhibited by PKC.Not tested[125]
Dorsal cochlear nucleus cartwheel cellsPostsynaptic
(soma and dendrites)
Control EPSP.IBTX-sensitive[113,175]
Retinal ganglion cellsPostsynaptic
(soma)
Regulate cell excitability.IBTX-sensitive[176]
A17 amacrine cellsPostsynaptic (varicosities)Limit Cav channel activation and GABA release.β2 subunit, IBTX-sensitive[22]
Abbreviations: AIS: axonal initial segment; IBTX: iberiotoxin; Pax: paxilline; Chtx: charybdotoxin; EPSP: excitatory postsynaptic potential; AP: action potential; AHP: afterhyperpolarization; DSE: depolarization-induced suppression of excitation.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Ancatén-González, C.; Segura, I.; Alvarado-Sánchez, R.; Chávez, A.E.; Latorre, R. Ca2+- and Voltage-Activated K+ (BK) Channels in the Nervous System: One Gene, a Myriad of Physiological Functions. Int. J. Mol. Sci. 2023, 24, 3407. https://doi.org/10.3390/ijms24043407

AMA Style

Ancatén-González C, Segura I, Alvarado-Sánchez R, Chávez AE, Latorre R. Ca2+- and Voltage-Activated K+ (BK) Channels in the Nervous System: One Gene, a Myriad of Physiological Functions. International Journal of Molecular Sciences. 2023; 24(4):3407. https://doi.org/10.3390/ijms24043407

Chicago/Turabian Style

Ancatén-González, Carlos, Ignacio Segura, Rosangelina Alvarado-Sánchez, Andrés E. Chávez, and Ramon Latorre. 2023. "Ca2+- and Voltage-Activated K+ (BK) Channels in the Nervous System: One Gene, a Myriad of Physiological Functions" International Journal of Molecular Sciences 24, no. 4: 3407. https://doi.org/10.3390/ijms24043407

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