Large conductance voltage- and calcium-activated potassium channels (BK channels) are extensively found throughout the central nervous system and play a crucial role in various neuronal functions. These channels are activated by a combination of cell membrane depolarisation and an increase in intracellular calcium concentration, provided by calcium sources located close to BK. In 2001, Isaacson and Murphy first demonstrated the coupling of BK channels with N-methyl-D-aspartate receptors (NMDAR) in olfactory bulb neurons. Since then, additional evidence has confirmed this functional coupling in other brain regions and highlighted its significance in neuronal function and pathophysiology. In this review, we explore the current understanding of these macrocomplexes in the brain, the molecular mechanisms behind their interactions and their potential roles in neurodevelopmental disorders, paving the way for new treatment strategies.

Large conductance voltage- and Ca2+- activated K+ channels (KCa1.1, BK, MaxiK or slo1) are widely expressed in diverse tissues and contribute to a myriad of specialised physiological functions [1]. Specific localisation within cell compartments is a key determinant of their physiological roles [2], greatly influenced by the association with auxiliary subunits and other proteins, including other ion channels or transporters [3]. In addition to their individual roles as regulators of cellular excitability or epithelial functions, growing evidence shows that the functional association between BK and various intracellular Ca2+ sources underlines a general mechanism regulating Ca2+ signalling in very diverse physiological contexts [4-8]. Here, we review how BK channels are involved in regulating synaptic transmission by their specific association with N-methyl-D-aspartate receptors (NMDAR). The unique characteristics associated with the function and regulation of BK channels set the background for very complex physiological contexts, which may provide the molecular basis for a large diversity of physiological outputs.

BK channels: physiological roles and significance in neuronal function

BK channels are expressed at the plasma membrane as homotetramers of α subunits encoded by the KCNMA1 (Slo1) gene [9]. Of note, functional expression of these channels has been shown in mitochondrial and nuclear membranes, although these will not be discussed in this review [10]. BK channels play a broad range of specialised physiological roles in a variety of excitable and non-excitable tissues, including muscle, kidney, gastrointestinal tract, salivary glands and bone (recently reviewed in González-Sanabria et al. [10]). In the central nervous system (CNS), they are largely expressed in soma, axons and synaptic terminals of cells from the olfactory system, neocortex, basal ganglia, hippocampus and thalamus [11-15]. BK function in excitable tissue has been related to shaping the action potential waveforms and influencing the firing frequency in various ways. Consistent with the established understanding of potassium channel function in excitability, several studies demonstrate that blocking BK channels with paxilline (PAX), charybdotoxin or tetraethylammonium (TEA) prolongs action potential repolarisation and reduces the amplitude of the fast-duration after-hyperpolarisation (AHP) [16-20]. The pharmacological or genetic removal of BK channels can either decrease or increase the frequency of both evoked and spontaneous firing [21-25]. The latter effect can be explained by BK decreasing the latency to bring voltage-activated Na+ and Ca2+ channels out of their inactivated state, allowing higher activation frequencies [26]. Additionally, the genetic deletion of the BKβ4 subunit in dentate gyrus (DG) neurons led to an increase in fast AHP amplitude, sharper action potentials and higher spike frequencies [27]. These findings suggest that BK channels are not purely excitatory or inhibitory; instead, they are dynamically regulated to control neuronal excitability, depending on the cellular context [28].

One key physiological feature of BK channel activity is the requirement of coincidental depolarisation and intracellular Ca2+ increase to be activated [29,30]. In fact, in many cell types, the activation of BK channels depends on the localised rise in Ca2+ levels reaching micromolar concentrations, which are significantly above the resting cytosolic values of 100 nm to 300 nm [31]. Thus, BK channels are commonly located in close proximity to other proteins acting as intracellular Ca2+ sources, associated functionally to form Ca2+ nano- or micro-domains (Figure 1). One such association which has been amply studied is that of BK with voltage-gated Ca2+ channels (VGCC). In neurons and smooth muscle cells, the synergistic membrane depolarisation from an action potential plus the Ca2+ entry through activated VGCC triggers nearby BK channels. The large outward K+ current helps repolarise the membrane, leading to the closure of the Ca2+ channels and ending the calcium signal [31,32]. As explained in the following sections, this negative feedback mechanism is not restricted to the functional coupling of BK to VGCC. It rather shows high versatility, including a large variety of Ca2+ sources [3,4]. Indeed, growing evidence shows the physiological relevance of this mechanism in a large variety of functions. These include action potential termination ([4,33] and see above), neurotransmitter release [8,34], control of circadian rhythm [35], smooth muscle contraction [36,37], cardiac sinoatrial node firing [38] and insulin secretion [39]. In line with their physiological roles, alterations in BK channels are associated with several genetically linked and acquired diseases (for recent reviews, see [6,40]).

Functional nanodomains and microdomains.

Figure 1:
Functional nanodomains and microdomains.

The biophysical features of BK channels are related with their functional association to different Ca2+ sources in various physiological settings. According to the distances between these different channels, such complexes (‘channelosomes’) can be defined as nanodomains (if the distance is 50 nm or lower) or microdomains (if the distance is higher than 50 nm). The association between BK and the Ca2+ sources is spatially and temporally regulated.

Figure 1:
Functional nanodomains and microdomains.

The biophysical features of BK channels are related with their functional association to different Ca2+ sources in various physiological settings. According to the distances between these different channels, such complexes (‘channelosomes’) can be defined as nanodomains (if the distance is 50 nm or lower) or microdomains (if the distance is higher than 50 nm). The association between BK and the Ca2+ sources is spatially and temporally regulated.

Close modal

To fully understand BK function and its connection to various calcium sources, it is important to consider the tissue-specific expression of KCNMA1 gene products through mechanisms such as alternative splicing [1,41,42] or post-translational modifications [43]. In addition, native BK channels are commonly found in conjunction with one of two classes of modulatory subunits, β(1-4) and γ(1-4), which differ in structure and function [44-49]. The presence of these subunits influences nearly all aspects of BK channel gating, including its kinetics, voltage sensitivity, Ca2+ responsiveness and pharmacological properties [1,49].

NMDA receptors as Ca2+ sources

NMDAR are heterotetrameric ligand-gated ion channels that belong to the family of glutamate-gated ion channels, also known as ionotropic glutamate receptors (iGluR), together with the α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid receptor (AMPAR) and kainate receptor [50,51]. The function of NMDAR, which in physiological conditions mediates the inflow of Na+ and Ca2+ and outflow K+, has been extensively studied [52]. Since NMDARs allow the graded entry of Ca2+ in the cell in response to ligand binding, they play a role in synaptic plasticity, learning, memory and other higher cognitive functions [53]. The physiological relevance of NMDAR is evidenced by the fact that malfunctioning of these receptors has been related to a variety of neurological and psychiatric disorders, including Alzheimer’s disease [54], Huntington’s disease [55], stroke and schizophrenia [53], as well as major depressive disorder [56].

Seven homologous NMDAR subunits have been described so far: the obligatory GluN1/NR1 subunit, four different GluN2/NR2 subunits (GluN2A, GluN2B, GluN2C and GluN2D) and two GluN3/NR3 subunits (GluN3A and GluN3B). GluN1 and GluN3 subunits contain a binding site for the co-agonist (glycine or D-serine), whereas GluN2 subunits exhibit an agonist (glutamate) binding site. NMDAR can assemble as two GluN1 and two GluN2 subunits (GluN1/GluN2), as well as di-heteromeric (GluN1/GluN3) or tri-heteromeric (GluN1/GluN2/GluN3) combinations, resulting in the vast functional diversity of NMDAR in the CNS [52]. The expression of these genes is spatially and temporally controlled, adding to NMDAR heterogeneity throughout the brain [53].

NMDARs are highly sensitive to glutamate, with a half-maximal effective concentration in the micromolar range, and are subject to voltage-dependent blockade by Mg2+ ions [57,58]. In addition, their slow gating kinetics [59] and significant Ca2+ permeability [60,61] enable postsynaptic NMDARs to detect and interpret the coinciding activity of both pre- and postsynaptic neurons. Specifically, presynaptic release of glutamate binds to the receptor, while postsynaptic depolarisation driven by AMPARs removes the Mg2+ block. This simultaneous occurrence activates NMDARs, allowing Ca2+ to enter through the channel, triggering signalling cascades that can modulate synaptic plasticity [53]. It is then inferred that any regulatory process affecting Ca2+ entry through NMDAR will modify neuronal plasticity and all related effects, as we will detail in the following sections.

The NMDAR-BK channelosome

The first evidence that Ca2+ entering through NMDARs could activate nearby Ca2+-activated potassium channels was reported in the 1980s in hippocampal neurons [62,63]. This mechanism was fully described in the olfactory bulb by Isaacson and Murphy [64], showing the activation of BK outward currents by the inflow of Ca2+ through glutamate-activated NMDAR. Zhang and colleagues showed that this functional association may be extended to numerous brain regions, as suggested biochemically by co-immunoprecipitation of BK and NMDAR in hippocampus, cortex, cerebellum, striatum and thalamus [65]. A role of these complexes has been reported in cortical neurons [15,66,67], hippocampal neurons [68,69], nucleus accumbens (NAc) [70], dorsal cochlear nucleus [71] and superficial dorsal horn (SDH) neurons of the thoracolumbar spinal cord [72]. Function of these associations has been tested by performing whole-cell patch-clamp recordings after applying glutamate or NMDA either towards the neuronal soma [65,67] or at dendritic locations [15,66]. Functional activation of BK by NMDAR in dendrites has been demonstrated in cortical layer 5 pyramidal neurons (L5PN) [15,66,73], as well as hippocampus CA3 and amygdala [74]. NMDAR-BK complexes have been described in both extrasynaptic and postsynaptic terminals [15,64-66,73].

The functional association of NMDAR and BK can be reproduced electrophysiologically in heterologous expression systems such as HEK cells, expanding the range of experimental approaches to understand the mechanisms underlying formation and function of these channelosomes [65,66]. As shown in Figure 2A, the whole-cell configuration of the patch-clamp technique can be used in transfected cells, rendering similar currents as those observed in native neurons [63-66]. Currents are recorded at different holding potentials, after the addition of NMDA or glutamate, in the absence of Mg2+. In those cells where NMDAR does not form functional nanodomains with BK, application of the NMDAR agonist produces inward currents (Figure 2A, ‘A-type’ recordings). In contrast, association of NMDAR to BK channels produces a slower voltage-dependent outward current at holding potentials more positive than −40 mV (Figure 2B, ‘B-type’ recordings). Further information can be inferred from the use of the inside-out excised patch configuration of the patch-clamp technique (Figure 2B). In this experimental setting, NMDARs are activated by including 200 μM NMDA and 10 μM glycine in the ‘extracellular’ pipette solution. The [Ca2+] in the bath can be controlled by the experimenter and is reported by the position of the BK G-V curve, which is left-shifted as [Ca2+]i is increased [75] (Figure 2B, black circles and coloured dashed lines). Interestingly, the recordings from patches co-expressing NMDAR and BK in a bath solution containing zero [Ca2+] produce G-V curves that are comparable with those obtained in the presence of 10 µM intracellular Ca2+ (Figure 2B, purple circles). The most feasible explanation is that activated NMDARs in the patch, closely located to co-expressed BK, supply Ca2+ from the pipette solution. One potential caveat of this approach is that overexpressing these two proteins might lead to artifactual co-localisation. To address this, it is important to conduct appropriate control experiments with other membrane proteins or channels to assess the specificity of multicomplex formation. In support of this, we have confirmed the specificity of the NMDAR-BK interaction by showing no association between BK and Ca2+-permeant AMPAR variants (GluA2(Q) [76]) or between NMDAR and Kv1.1 (a voltage-gated potassium channel), under similar experimental conditions (unpublished data). In any case, and similarly to what would be done in other studies using heterologous expression systems, conclusions must take into account the limitations of this experimental model and combine it with studies in native systems. This experimental approach can be used, for instance, to evaluate the effect of different subunit combinations on the functional coupling [66]. Additionally, it constitutes a useful tool to assess the effect of mutations on BK or NMDAR affecting the functional coupling [77], as well as other remaining questions regarding the structural, biophysical and pharmacological properties of these multiprotein complexes.

NMDAR-BK nanodomains can be reconstituted in heterologous expression systems.

Figure 2:
NMDAR-BK nanodomains can be reconstituted in heterologous expression systems.

(A) Top scheme represents the cell being recorded using the whole-cell configuration of the patch-clamp technique. Bottom, representative currents recorded at different holding potentials, after the addition of NMDA or glutamate, in Mg2+-free solutions (see [64-66] for experimental details). ‘A-type’ cells (left, red traces) lack NMDAR-BK macrocomplexes, whereas ‘B-type’ cells (right, blue traces) express NMDAR-BK functional associations. (B) Top, scheme representing the inside-out excised patch configuration of the patch-clamp technique. Bottom, G-V curves obtained from patches expressing BK channels alone (black circles and coloured dashed lines) or co-expressed with NMDAR (purple circles). For experimental details, see main text and Gómez et al. [66]. Part of the data shown has been published in Gómez et al. [66].

Figure 2:
NMDAR-BK nanodomains can be reconstituted in heterologous expression systems.

(A) Top scheme represents the cell being recorded using the whole-cell configuration of the patch-clamp technique. Bottom, representative currents recorded at different holding potentials, after the addition of NMDA or glutamate, in Mg2+-free solutions (see [64-66] for experimental details). ‘A-type’ cells (left, red traces) lack NMDAR-BK macrocomplexes, whereas ‘B-type’ cells (right, blue traces) express NMDAR-BK functional associations. (B) Top, scheme representing the inside-out excised patch configuration of the patch-clamp technique. Bottom, G-V curves obtained from patches expressing BK channels alone (black circles and coloured dashed lines) or co-expressed with NMDAR (purple circles). For experimental details, see main text and Gómez et al. [66]. Part of the data shown has been published in Gómez et al. [66].

Close modal

Molecular composition of the NMDAR-BK macrocomplexes

It has been proposed that the interaction between NMDAR and BK occurs via intracellular interactions between the GluN1 and BKα subunits [65]. Zhang et al. [65] demonstrated in vitro the interaction of the isolated GluN1 cytosolic regions with a synthesised peptide of the BKα S0–S1 loop region. In addition, these authors showed that this interaction was competitively diminished by a synthesised peptide from BKα S0–S1 loop [65]. These findings may suggest that, if the main interaction occurs between the obligatory GluN1 subunit and BKα, the role of the different GluN2 subunits would not affect the functional association. However, this aspect has not been fully addressed. Even though the proposed GluN1-BKα interactions may be required to form the macrocomplexes, the presence of different GluN2 regulatory subunits may introduce diversity in the biophysical properties of the nanodomains and thus in their physiological roles, such as the fine-tuning of synaptic plasticity (see below). In line with this idea, in vivo experiments using the excised inside-out patch technique (Figure 2B) with solutions containing physiological concentrations of Na+, GluN1/GluN2B NMDARs produced a larger leftward shift in the BK G-V activation curve than GluN1/GluN2A. This observation correlated with data from basal dendrites of barrel cortex L5PNs (BC-L5PNs), where the specific blockade of GluN2B-containing NMDARs resulted in a larger reduction of the NMDA-evoked outward currents [66]. Taking into account that the subunit composition of NMDARs is dynamic and varies quickly throughout development, influenced by neuronal activity or sensory experiences, even at adult synapses [53], it is tempting to speculate that differences in NMDAR composition may impact the function of NMDAR-BK complexes and therefore their effects on synaptic function.

Recordings from two distinct subpopulations of BC-L5PNs which were differentiated by the presence or absence of NMDAR-BK complexes in basal dendrites showed comparable NMDAR current characteristics, indicating a similar distribution of GluN2 subunits. Interestingly, both neuronal types showed BK channel activity. The easiest explanation for the presence or absence of NMDAR-BK complexes would be the existence of specific mechanisms targeting the channels to dendritic compartments. This could be possibly achieved by engaging specific scaffolding proteins, including the receptor for activated C kinase 1 (RACK1) and caveolin-1, both known to bind the GluN2B NMDAR subunit [78,79] and BK channels [80,81]. To address this question, protein profiling either in heterologous expression systems or native neurons would be a powerful approach to screen for candidate proteins forming the NMDAR-BK interactome [82].

Interestingly, the interaction between BK and NMDAR can be altered by GluN2B mutations related to disease. These observations show that the GluN2B mutation V618G, located in the transmembrane domain [83-85], results in weaker functional coupling, which is independent of Ca2+ permeation or NMDAR expression levels [77]. Strikingly, NMDAR-BK cluster quantification using superresolution microscopy shows that the size of the macrocomplexes is significantly smaller. In addition, the proportion of NMDAR and BK particles is altered in the presence of the mutation. These novel findings suggest that, even though NDMAR and BK may form complexes in different conditions, the functional characteristics of NMDAR associations are affected by the size of the clusters and by the relative number of NMDAR and BK conforming to the nanodomain [77]. It may also be inferred that the formation of NMDAR-BK macrocomplexes should not be exclusively ascribed to the proposed GluN1-BK interactions [65]. It is tempting to speculate that these alterations in the functional NMDAR-BK associations, even if partial, may be related to the pathophysiological effects of the NMDAR-GluN2B mutations. Determining the impact on NMDAR-BK function of other clinically relevant mutations in GluN2A and GluN2B subunits requires further investigation. Analogously, the effects of BK variants associated with CNS pathologies should be explored [5,40,86]. Undoubtedly, structural determination of the NMDAR-BK complexes (including the abovementioned pathological variants) would dramatically advance our understanding about the molecular interactions underlying the function and assembly of these multichannel associations.

Finally, it must be taken into consideration that the molecular mechanisms underlying NMDAR-BK interactions and function might be influenced by other factors related to the functional complexity of BK channels, including regulation by multiple auxiliary subunits (mainly β2 and β4 in the nervous system), alternative splicing variants or post-translational modifications of BKα [1]. The function of NMDAR-BK macrocomplexes remains to be studied in these variable physiological scenarios.

Physiological roles of NMDAR-BK functional associations

All the abovementioned evidence leads to the conclusion that NMDAR-BK complexes constitute a general control mechanism in the CNS, contributing to neuronal function in different brain regions [15,62-66,74]. These interactions occur both in somatic [63-65] and dendritic locations [15,66], where they may play different roles. At the soma, the NMDAR-dependent activation of BK channels provides a mechanism to regulate action potential shape and neuronal excitability, separately from dendritic input. It is worth noting that, in this subcellular context, the diverse roles of BK must be taken into account (see above), adding a layer of complexity to the proposed model. In dendrites, the coupling of NMDAR to BK may provide a negative feedback mechanism regulating synaptic transmission and plasticity (Figure 3) [15,65,66,72]. Two different physiological settings may be distinguished by the absence in dendrites of NMDAR-BK functional nanodomains (Figure 3, left, ‘A-type’ neurons) or the presence of these multichannel complexes (Figure 3, right, ‘B-type’ neurons). These neuronal types are characterised by distinct properties of their NMDA-activated currents, with important implications in their synaptic transmission and plasticity properties. In A-type neurons, the entrance of Ca2+ through NMDAR, which is activated by coincident glutamate binding and AMPAR-mediated depolarisation, as described above, has been proposed to activate the Ca2+/calmodulin-dependent protein kinase II (CamKII), which subsequently facilitates the trafficking and stabilisation of AMPARs at synapses, inducing long-term potentiation (LTP) (for extended reviews, see [53,87]). As previously described for heterologous expression systems (Figure 2), A-type neurons are characterised by NMDAR-like inward currents after addition of NMDA or glutamate to dendrites (Figure 3, bottom left panel). In these neurons, recordings of synaptically evoked post-synaptic potentials (PSPs; traces on top of the figure) showed no significant changes after blockade of BK channels with PAX (Figure 3, left, upper traces). Conversely, in B-type neurons the association of NMDAR to BK channels provides a negative feedback mechanism by which Ca2+ entry through activated NMDAR opens adjacent BK channels, allowing K+ to flow outside the cell (Figure 3, right bottom traces). The resultant membrane hyperpolarisation restores the voltage-dependent Mg2+ block of NMDARs, abolishing Ca2+ entry and augmenting the threshold for LTP [66]. A possible mechanism to explain the reduced LTP levels would be that the reduced Ca2+ entry results in subthreshold activation of CamKII, therefore blunting the associated increase in AMPAR trafficking. B-type neurons are characterised by NMDAR-like AP5-sensitive inward currents followed by slow outward currents (Figures 1 and 3), which are blocked by PAX (Figure 3, right; black traces at the bottom). Both inward and outward currents are abolished by AP5 (Figure 3, right; light blue traces at the bottom), demonstrating that BK activation is driven by Ca2+ entering through NMDARs [64-66]. Consistent with the negative feedback effect played by BK, the blockade of these channels with PAX causes an increase in the synaptically evoked post-synaptic potentials (PSPs) (Figure 3, right; upper recordings) [65,66]. A similar effect of PAX on PSPs was reported in inhibitory interneurons from lamina II of the rat spinal dorsal horn [72]. Interestingly, blocking of BK mediated by cholinergic activation in cartwheel cells of the dorsal cochlear nucleus results in enhanced excitatory post-synaptic potentials (EPSPs) and spine Ca transients [71]. In lateral amygdala (LA), reduced expression of BK channels induced by acute stress produced an increase in the evoked NMDA receptor-mediated EPSPs at the thalamo-LA synapses [88]. In BC-L5PN, it has been demonstrated that the presence of NMDAR-BK functional coupling results in reduced synaptic transmission and a higher threshold for the induction of LTP [66]. The selective plasticity attenuation exerted by NMDAR-BK macrocomplexes is restricted to the basal dendrites of these neurons since stimulation of apical dendrites did not produce any effect [15,66]. Notably, a related effect has been observed in NAc, where BK channels have been involved in LTP inhibition by ethanol (EtOH) [70]. In NAc medium spiny neurons, BK function potentiated by EtOH would more effectively counteract NMDAR activity, therefore amplifying the effect of EtOH on NMDAR [70]. Reduced synaptic transmission associated with NMDAR-BK function has been also observed in mature DG granule cells [65]. It has been proposed that this dendritic regulatory mechanism could serve to interpret the quantity and frequency of afferent synaptic inputs by selectively reducing synaptic plasticity and introducing input-specific synaptic diversity [15,66]. This process has been proposed to be further regulated by the spine structure [15].

NMDAR-BK channelosomes as regulators of synaptic transmission and plasticity.

Figure 3:
NMDAR-BK channelosomes as regulators of synaptic transmission and plasticity.

Cartoon representing the biophysical and physiological features of neurons lacking dendritic NMDAR-BK multichannel complexes (‘A-type’ neurons, left, red) versus neurons containing these functional nanodomains (‘B-type’ neurons, right, blue). Inset on the right describes the molecular components shown in the cartoon. In A-type neurons (left), the entrance of Ca2+ through NMDAR [1] activates Ca2+/calmodulin-dependent protein kinase II, facilitating trafficking and stabilisation of α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid receptors at synapses [2], therefore inducing long-term potentiation (LTP [3]). Red current traces at the bottom represent typical responses to glutamate. Red traces at the top represent synaptically evoked PSPs before and after treatment with PAX and PAX+AP5. In B-type neurons (right), Ca2+ entry through activated NMDARs [1] opens adjacent BK channels [2] allowing K+ to flow outside the cell, leading to membrane hyperpolarisation (-Vm), which restores the voltage-dependent Mg2+block of NMDARs [3], abolishing Ca2+ entry and augmenting the threshold for LTP. Blue current traces at the bottom represent typical responses to glutamate and its response to different blockers (black, PAX; light blue, PAX+AP5). Blue traces at the top represent synaptically evoked PSPs before and after treatment with PAX and PAX+AP5. Part of the data shown has been published in Gómez et al. [66]. See details in main text and Gómez et al. [66].

Figure 3:
NMDAR-BK channelosomes as regulators of synaptic transmission and plasticity.

Cartoon representing the biophysical and physiological features of neurons lacking dendritic NMDAR-BK multichannel complexes (‘A-type’ neurons, left, red) versus neurons containing these functional nanodomains (‘B-type’ neurons, right, blue). Inset on the right describes the molecular components shown in the cartoon. In A-type neurons (left), the entrance of Ca2+ through NMDAR [1] activates Ca2+/calmodulin-dependent protein kinase II, facilitating trafficking and stabilisation of α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid receptors at synapses [2], therefore inducing long-term potentiation (LTP [3]). Red current traces at the bottom represent typical responses to glutamate. Red traces at the top represent synaptically evoked PSPs before and after treatment with PAX and PAX+AP5. In B-type neurons (right), Ca2+ entry through activated NMDARs [1] opens adjacent BK channels [2] allowing K+ to flow outside the cell, leading to membrane hyperpolarisation (-Vm), which restores the voltage-dependent Mg2+block of NMDARs [3], abolishing Ca2+ entry and augmenting the threshold for LTP. Blue current traces at the bottom represent typical responses to glutamate and its response to different blockers (black, PAX; light blue, PAX+AP5). Blue traces at the top represent synaptically evoked PSPs before and after treatment with PAX and PAX+AP5. Part of the data shown has been published in Gómez et al. [66]. See details in main text and Gómez et al. [66].

Close modal

The argument for functional interactions between NMDARs and BK channels is further bolstered by a recent study by Pham et al. [89], which revealed that BK channels facilitate NMDAR-mediated Ca2+ entry in hippocampal neurons, with the absence of BK activity impairing hippocampal synaptic plasticity [89]. Notably, signalling within the NMDAR-BK macrocomplexes appears to be bidirectional, as the functional components mutually influence each other. These results are relevant especially in light of findings that both gain- and loss-of-function mutations in BK have been associated with developmental delays and intellectual disabilities [90], potentially linked to disruptions in synaptic plasticity.

The influence of NMDAR-BK complexes in regulating synaptic plasticity in other brain regions such as the DG, hippocampal CA3 neurons or basolateral amygdala, where the presence of these associations has been shown [15,63-66,74], remains to be fully studied. As mentioned above, the role of NMDAR subunit composition and BK complexity (regulatory subunits, splicing variants or post-translational modifications) must be taken into account when addressing these relevant questions. For example, Ji et al. [70] found that mice lacking the β4 subunit showed dramatically accelerated EPSPs and reduced spike-timing dependent long-term potentiation (tLTP) [70] in the NAc, suggesting a role of these regulatory subunits in the function of NMDAR-BK complexes.

The relevance of the NMDAR-BK association has been recently demonstrated in some pathological scenarios. In the rat spinal cord, it has been proposed that BK channels interact with NMDAR to regulate visceral pain transmission and visceral hypersensitivity in a model of irritable bowel syndrome [72]. In a mouse model of Fragile X syndrome, BK-dependent synaptic integration [73] and NMDAR-BK coupling [74] are significantly altered, indicating that the integrity of this mechanism is essential in the healthy brain.

An interesting aspect that deserves further attention is the differential distribution of A-type and B-type neurons in specific brain areas. BC-L5PNs show two subpopulations with A-type (60%) or B-type (40%) features [66]. In contrast, all neurons in the CA3 hippocampus or basolateral amygdala show B-type properties [74]. The physiological implications of these neuronal population distributions on synaptic and network functions warrant further attention, as well as the exploration of other brain areas with this new functional focus.

Finally, a key question remains about the physiological roles that NMDAR-BK associations may play in different neuronal locations. In the adult brain, the majority of NMDARs located at synaptic sites are di-heteromeric GluN1-GluN2A or tri-heteromeric GluN1-GluN2A-GluN2B receptors. On the other hand, peri and extrasynaptic sites mainly contain GluN1-GluN2B receptors. In addition, GluN2C- and GluN2D-containing NMDARs can participate in synaptic transmission in some brain areas [53]. In all cases, NMDARs have been shown to be highly mobile, exchanging between synaptic and extrasynaptic sites [91]. This fact raises a very interesting question regarding the dynamics of NMDAR-BK associations. Can they form in response to neuronal activity? To this end, experiments using optogenetics could be performed to combine selected stimulation of brain areas with electrophysiological experiments discussed above. Is it possible that NMDAR-BK associations are differently shaped during development or regulated by different components or signalling pathways depending on the pathophysiological context? For instance, it has been proposed that interleukin-1β (IL-1β) may increase excitability in dissociated hippocampal neurons via regulation of NMDAR-BK function [68,69]. NMDAR-BK functional association has been suggested to be involved in the neuroprotective role of prostaglandin E2 (EP2) receptor-mediated signalling pathways in cortical neurons [67], cholinergic signalling in the dorsal cochlear nucleus [71] and inhibition of spinal opioid release [92]. Furthermore, presynaptic NMDARs have been described, which may form macrocomplexes with BK, similar to the well-studied presynaptic BK-VGCC associations [31]. Finally, the putative association of BK with non-neuronal NMDARs in astrocytes and oligodendrocytes remains largely unexplored. The role of NMDAR-BK associations in regulating synaptic transmission and plasticity seems to be attributable to their expression in postsynaptic sites, both in L5PN and dentate granular cells [15,65,66,74]. However, the expression of BK has been reported in most sites where NMDARs are expressed [53,86], suggesting that novel regulatory roles of NMDAR-BK complexes may be unveiled in the future.

Conclusions

NMDAR-BK complexes have been known for more than 20 years, but their role(s) remain largely unexplored. Growing evidence demonstrates that these functional associations play relevant roles in the CNS, with important implications for synaptic pathologies. Further study is required to understand the molecular basis of this interaction, the location and function of these channelosomes in different CNS areas and potential roles in regulating NMDAR-dependent neuronal processes.

Perspectives

  • N-methyl-D-aspartate receptors (NMDAR)-BK multichannel complexes constitute a regulatory mechanism of excitability and synaptic function, which is present in many neuronal types and may have been overseen in many physiological settings.

  • Current roles of NMDAR-BK include the regulation of synaptic transmission and plasticity, serving as high-pass filters for incoming synaptic inputs.

  • Future directions should include the characterisation of NMDAR-BK complexes in different pathophysiological settings, unveiling novel targets for neurological diseases and synaptopathies.

The authors declare that there are no competing interests associated with the manuscript.

Grant PID2021-128668OB-I00 funded by MICIU/AEI/10.13039/501100011033 and by 'ERDF/EU' (To TG); Grant PRE2019-089248 funded MICIU/AEI/10.13039/501100011033 and 'ESF Investing in your future' (To RML); Grant TESIS2021010099 from Gobierno de Canarias (To ARC).

This article has been published open access under our Subscribe to Open programme, made possible through the support of our subscribing institutions, learn more here: https://portlandpress.com/pages/open_access_options_and_prices#conditional

R.-M.L., A.R.-C., D.B.-M. and T.G. conceived the work and wrote the manuscript.

We are grateful to Diego Alvarez de la Rosa for useful discussions and comments on the manuscript.

AHP

after-hyperpolarisation

AMPAR

α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid receptor

BC-L5PNs

barrel cortex L5PNs

BK channels

large conductance voltage- and calcium-activated potassium channels

CNS

central nervous system

CamKII

calmodulin-dependent protein kinase II

DG

dentate gyrus

EP2

prostaglandin E2

EPSPs

excitatory post-synaptic potentials

IL

interleukin

LA

lateral amygdala

L5PN

layer 5 pyramidal neurons

LTP

long-term potentiation

NAc

nucleus accumbens

NMDAR

N-methyl-D-aspartate receptors

PAX

paxilline

PSPs

post-synaptic potentials

RACK1

receptor for activated C kinase 1

SDH

superficial dorsal horn

VGCC

voltage-gated Ca2+ channels

iGluR

ionotropic glutamate receptors

tLTP

spike-timing dependent long-term potentiation

1
Latorre
,
R.
,
Castillo
,
K.
,
Carrasquel-Ursulaez
,
W.
,
Sepulveda
,
R.V.
,
Gonzalez-Nilo
,
F.
,
Gonzalez
,
C.
et al.
(
2017
)
Molecular Determinants of BK Channel Functional Diversity and Functioning
.
Physiol. Rev.
97
,
39
87
https://doi.org/10.1152/physrev.00001.2016
2
Bock
,
T.
and
Stuart
,
G.J
. (
2016
)
The Impact of BK Channels on Cellular Excitability Depends on their Subcellular Location
.
Front. Cell. Neurosci.
10
, 206 https://doi.org/10.3389/fncel.2016.00206
3
Gonzalez-Hernandez
,
AJ.
,
Kshatri
,
A.
and
Giraldez
,
T
. (
2023
)
Calcium-activated potassium channels. 2023. In: Calcium Signals [Internet], IOP Publishing
. https://doi.org/10.1088/978-0-7503-2009-2ch6
4
Shah
,
K.R.
,
Guan
,
X.
and
Yan
,
J
. (
2021
)
Structural and Functional Coupling of Calcium-Activated BK Channels and Calcium-Permeable Channels Within Nanodomain Signaling Complexes
.
Front. Physiol.
12
, 796540 https://doi.org/10.3389/fphys.2021.796540
5
Kshatri
,
A.S.
,
Gonzalez-Hernandez
,
A.J.
and
Giraldez
,
T
. (
2018
)
Physiological Roles and Therapeutic Potential of Ca2+ Activated Potassium Channels in the Nervous System
.
Front. Mol. Neurosci.
11
, 258 https://doi.org/10.3389/fnmol.2018.00258
6
Echeverría
,
F.
,
Gonzalez-Sanabria
,
N.
,
Alvarado-Sanchez
,
R.
,
Fernández
,
M.
,
Castillo
,
K.
and
Latorre
,
R
. (
2024
)
Large conductance voltage-and calcium-activated K+ (BK) channel in health and disease
.
Front. Pharmacol.
15
, 1373507 https://doi.org/10.3389/fphar.2024.1373507
7
Gonzalez-Perez
,
V.
,
Martinez-Espinosa
,
P.L.
,
Sala-Rabanal
,
M.
,
Bharadwaj
,
N.
,
Xia
,
X.-M.
,
Chen
,
A.C.
et al.
(
2021
)
Goblet cell LRRC26 regulates BK channel activation and protects against colitis in mice
.
Proc. Natl. Acad. Sci. U.S.A.
118
, e2019149118 https://doi.org/10.1073/pnas.2019149118
8
Lingle
,
C.J.
,
Solaro
,
C.R.
,
Prakriya
,
M.
and
Ding
,
J.P
. (
1996
)
Calcium-activated potassium channels in adrenal chromaffin cells
.
Ion Channels
4
,
261
301
https://doi.org/10.1007/978-1-4899-1775-1_7
9
Cui
,
J.
,
Yang
,
H.
and
Lee
,
U.S
. (
2009
)
Molecular mechanisms of BK channel activation
.
Cell. Mol. Life Sci.
66
,
852
875
https://doi.org/10.1007/s00018-008-8609-x
10
González-Sanabria
,
N.
,
Echeverría
,
F.
,
Segura
,
I.
,
Alvarado-Sánchez
,
R.
and
Latorre
,
R
. (
2021
)
BK in double-membrane organelles: A biophysical, pharmacological, and functional survey
.
Front. Physiol.
12
, 761474 https://doi.org/10.3389/fphys.2021.761474
11
Sausbier
,
U.
,
Sausbier
,
M.
,
Sailer
,
C.A.
,
Arntz
,
C.
,
Knaus
,
H.-G.
,
Neuhuber
,
W.
et al.
(
2006
)
Ca2+-activated K+ channels of the BK-type in the mouse brain
.
Histoch. Cell Biol.
125
,
725
741
https://doi.org/10.1007/s00418-005-0124-7
12
Kshatri
,
A.S.
,
Gonzalez-Hernandez
,
A.
and
Giraldez
,
T
. (
2018
)
Physiological Roles and Therapeutic Potential of Ca2+ Activated Potassium Channels in the Nervous System
.
Front. Mol. Neurosci.
11
, 258 https://doi.org/10.3389/fnmol.2018.00258
13
Trimmer
,
J.S
. (
2015
)
Subcellular localization of K+ channels in mammalian brain neurons: remarkable precision in the midst of extraordinary complexity
.
Neuron
85
,
238
256
https://doi.org/10.1016/j.neuron.2014.12.042
14
Wang
,
B.
,
Jaffe
,
D.B.
and
Brenner
,
R
. (
2014
)
Current understanding of iberiotoxin-resistant BK channels in the nervous system
.
Front. Physiol.
5
, 382 https://doi.org/10.3389/fphys.2014.00382
15
Tazerart
,
S.
,
Blanchard
,
M.G.
,
Miranda-Rottmann
,
S.
,
Mitchell
,
D.E.
,
Navea Pina
,
B.
,
Thomas
,
C.I.
et al.
(
2022
)
Selective activation of BK channels in small-headed dendritic spines suppresses excitatory postsynaptic potentials
.
J. Physiol. (Lond.)
600
,
2165
2187
https://doi.org/10.1113/JP282303
16
Adams
,
P.R.
,
Constanti
,
A.
,
Brown
,
D.A.
and
Clark
,
R.B
. (
1982
)
Intracellular Ca2+ activates a fast voltage-sensitive K+ current in vertebrate sympathetic neurones
.
Nat. New Biol.
296
,
746
749
https://doi.org/10.1038/296746a0
17
Storm
,
J.F
. (
1987
)
Action potential repolarization and a fast after-hyperpolarization in rat hippocampal pyramidal cells
.
J. Physiol. (Lond.)
385
,
733
759
https://doi.org/10.1113/jphysiol.1987.sp016517
18
Sah
,
P.
and
McLachlan
,
E.M
. (
1992
)
Potassium currents contributing to action potential repolarization and the afterhyperpolarization in rat vagal motoneurons
.
J. Neurophysiol.
68
,
1834
1841
https://doi.org/10.1152/jn.1992.68.5.1834
19
Zhang
,
L.
and
McBain
,
C.J
. (
1995
)
Potassium conductances underlying repolarization and after-hyperpolarization in rat CA1 hippocampal interneurones
.
J. Physiol. (Lond.)
488 ( Pt 3)
,
661
672
https://doi.org/10.1113/jphysiol.1995.sp020998
20
Faber
,
E.S.L.
and
Sah
,
P
. (
2002
)
Physiological role of calcium-activated potassium currents in the rat lateral amygdala
.
J. Neurosci.
22
,
1618
1628
https://doi.org/10.1523/JNEUROSCI.22-05-01618.2002
21
Gu
,
N.
,
Vervaeke
,
K.
and
Storm
,
J.F
. (
2007
)
BK potassium channels facilitate high-frequency firing and cause early spike frequency adaptation in rat CA1 hippocampal pyramidal cells
.
J. Physiol. (Lond.)
580
,
859
882
https://doi.org/10.1113/jphysiol.2006.126367
22
Bielefeldt
,
K.
and
Jackson
,
M.B
. (
1993
)
A calcium-activated potassium channel causes frequency-dependent action-potential failures in A mammalian nerve terminal
.
J. Neurophysiol.
70
,
284
298
https://doi.org/10.1152/jn.1993.70.1.284
23
Jin
,
W.
,
Sugaya
,
A.
,
Tsuda
,
T.
,
Ohguchi
,
H.
and
Sugaya
,
E
. (
2000
)
Relationship between large conductance calcium-activated potassium channel and bursting activity
.
Brain Res.
860
,
21
28
https://doi.org/10.1016/s0006-8993(00)01943-0
24
Gittis
,
A.H.
,
Moghadam
,
S.H.
du Lac
,
S
2010
)
Mechanisms of sustained high firing rates in two classes of vestibular nucleus neurons: differential contributions of resurgent Na, Kv3, and BK currents
.
J. Neurophysiol.
104
,
1625
1634
https://doi.org/10.1152/jn.00378.2010
25
Pérez
,
G.J.
,
Desai
,
M.
,
Anderson
,
S.
and
Scornik
,
F.S
. (
2013
)
Large-conductance calcium-activated potassium current modulates excitability in isolated canine intracardiac neurons
.
Am. J. Physiol., Cell Physiol.
304
,
C280
6
https://doi.org/10.1152/ajpcell.00148.2012
26
Sausbier
,
M.
,
Hu
,
H.
,
Arntz
,
C.
,
Feil
,
S.
,
Kamm
,
S.
,
Adelsberger
,
H.
et al.
(
2004
)
Cerebellar ataxia and Purkinje cell dysfunction caused by Ca2+-activated K+ channel deficiency
.
Proc. Natl. Acad. Sci. U.S.A.
101
,
9474
9478
https://doi.org/10.1073/pnas.0401702101
27
Brenner
,
R.
,
Chen
,
Q.H.
,
Vilaythong
,
A.
,
Toney
,
G.M.
,
Noebels
,
J.L.
and
Aldrich
,
R.W
. (
2005
)
BK channel beta4 subunit reduces dentate gyrus excitability and protects against temporal lobe seizures
.
Nat. Neurosci.
8
,
1752
1759
https://doi.org/10.1038/nn1573
28
Contet
,
C.
,
Goulding
,
S.P.
,
Kuljis
,
D.A.
and
Barth
,
A.L
. (
2016
)
BK Channels in the Central Nervous System
.
Int. Rev. Neurobiol.
128
,
281
342
https://doi.org/10.1016/bs.irn.2016.04.001
29
Pallotta
,
B.S.
,
Magleby
,
K.L.
and
Barrett
,
J.N
. (
1981
)
Single channel recordings of Ca2+-activated K+ currents in rat muscle cell culture
.
Nat. New Biol.
293
,
471
474
https://doi.org/10.1038/293471a0
30
Marty
,
A
. (
1981
)
Ca-dependent K channels with large unitary conductance in chromaffin cell membranes
.
Nat. New Biol.
291
,
497
500
https://doi.org/10.1038/291497a0
31
Fakler
,
B.
and
Adelman
,
J.P
. (
2008
)
Control of K(Ca) channels by calcium nano/microdomains
.
Neuron
59
,
873
881
https://doi.org/10.1016/j.neuron.2008.09.001
32
Griguoli
,
M.
,
Sgritta
,
M.
and
Cherubini
,
E
. (
2016
)
Presynaptic BK channels control transmitter release: physiological relevance and potential therapeutic implications
.
J. Physiol. (Lond.)
594
,
3489
3500
https://doi.org/10.1113/JP271841
33
Montgomery
,
J.R.
and
Meredith
,
A.L
. (
2012
)
Genetic activation of BK currents in vivo generates bidirectional effects on neuronal excitability
.
Proc. Natl. Acad. Sci. U.S.A.
109
,
18997
19002
https://doi.org/10.1073/pnas.1205573109
34
Robitaille
,
R.
,
Garcia
,
M.L.
,
Kaczorowski
,
G.J.
and
Charlton
,
M.P
. (
1993
)
Functional colocalization of calcium and calcium-gated potassium channels in control of transmitter release
.
Neuron
11
,
645
655
https://doi.org/10.1016/0896-6273(93)90076-4
35
Meredith
,
A.L.
,
Wiler
,
S.W.
,
Miller
,
B.H.
,
Takahashi
,
J.S.
,
Fodor
,
A.A.
,
Ruby
,
N.F.
et al.
(
2006
)
BK calcium-activated potassium channels regulate circadian behavioral rhythms and pacemaker output
.
Nat. Neurosci.
9
,
1041
1049
https://doi.org/10.1038/nn1740
36
Nelson
,
M.T.
,
Cheng
,
H.
,
Rubart
,
M.
,
Santana
,
L.F.
,
Bonev
,
A.D.
,
Knot
,
H.J.
et al.
(
1995
)
Relaxation of arterial smooth muscle by calcium sparks
.
Science
270
,
633
637
https://doi.org/10.1126/science.270.5236.633
37
Meredith
,
A.L.
,
Thorneloe
,
K.S.
,
Werner
,
M.E.
,
Nelson
,
M.T.
and
Aldrich
,
R.W
. (
2004
)
Overactive bladder and incontinence in the absence of the BK large conductance Ca2+-activated K+ channel
.
J. Biol. Chem.
279
,
36746
36752
https://doi.org/10.1074/jbc.M405621200
38
Lai
,
M.H.
,
Wu
,
Y.
,
Gao
,
Z.
,
Anderson
,
M.E.
,
Dalziel
,
J.E.
and
Meredith
,
A.L
. (
2014
)
BK channels regulate sinoatrial node firing rate and cardiac pacing in vivo
.
Am. J. Physiol. Heart Circ. Physiol.
307
,
H1327
38
https://doi.org/10.1152/ajpheart.00354.2014
39
Houamed
,
K.M.
,
Sweet
,
I.R.
and
Satin
,
L.S
. (
2010
)
BK channels mediate a novel ionic mechanism that regulates glucose-dependent electrical activity and insulin secretion in mouse pancreatic β-cells
.
J. Physiol. (Lond.)
588
,
3511
3523
https://doi.org/10.1113/jphysiol.2009.184341
40
Meredith
,
A.L
. (
2024
)
BK Channelopathies and KCNMA1-Linked disease models
.
Annu. Rev. Physiol.
86
https://doi.org/10.1146/annurev-physiol-030323-042845
41
Tseng-Crank
,
J.
,
Foster
,
C.D.
,
Krause
,
J.D.
,
Mertz
,
R.
,
Godinot
,
N.
,
DiChiara
,
T.J.
et al.
(
1994
)
Cloning, expression, and distribution of functionally distinct Ca(2+)-activated K+ channel isoforms from human brain
.
Neuron
13
,
1315
1330
https://doi.org/10.1016/0896-6273(94)90418-9
42
Shipston
,
M.J
. (
2001
)
Alternative splicing of potassium channels: a dynamic switch of cellular excitability
.
Trends Cell Biol.
11
,
353
358
https://doi.org/10.1016/s0962-8924(01)02068-2
43
Shipston
,
M.J.
and
Tian
,
L
. (
2016
)
Posttranscriptional and Posttranslational Regulation of BK Channels
.
Int. Rev. Neurobiol.
128
,
91
126
https://doi.org/10.1016/bs.irn.2016.02.012
44
Yan
,
J.
and
Aldrich
,
R.W
. (
2010
)
LRRC26 auxiliary protein allows BK channel activation at resting voltage without calcium
.
Nat. New Biol.
466
,
513
516
https://doi.org/10.1038/nature09162
45
Yan
,
J.
and
Aldrich
,
R.W
. (
2012
)
BK potassium channel modulation by leucine-rich repeat-containing proteins
.
Proc. Natl. Acad. Sci. U.S.A.
109
,
7917
7922
https://doi.org/10.1073/pnas.1205435109
46
Redhardt
,
M.
,
Raunser
,
S.
and
Raisch
,
T
. (
2024
)
Cryo-EM structure of the Slo1 potassium channel with the auxiliary γ1 subunit suggests a mechanism for depolarization-independent activation
.
FEBS Lett.
598
,
875
888
https://doi.org/10.1002/1873-3468.14863
47
Yamanouchi
,
D.
,
Kasuya
,
G.
,
Nakajo
,
K.
,
Kise
,
Y.
and
Nureki
,
O
. (
2023
)
Dual allosteric modulation of voltage and calcium sensitivities of the Slo1-LRRC channel complex
.
Mol. Cell.
83
,
4555
4569
https://doi.org/10.1016/j.molcel.2023.11.005
48
Brenner
,
R.
,
Jegla
,
T.J.
,
Wickenden
,
A.
,
Liu
,
Y.
and
Aldrich
,
R.W
. (
2000
)
Cloning and functional characterization of novel large conductance calcium-activated potassium channel beta subunits, hKCNMB3 and hKCNMB4
.
J. Biol. Chem.
275
,
6453
6461
https://doi.org/10.1074/jbc.275.9.6453
49
Gonzalez-Perez
,
V.
and
Lingle
,
C.J
. (
2019
)
Regulation of BK Channels by Beta and Gamma Subunits
.
Annu. Rev. Physiol.
81
,
113
137
https://doi.org/10.1146/annurev-physiol-022516-034038
50
Hansen
,
K.B.
,
Wollmuth
,
L.P.
,
Bowie
,
D.
,
Furukawa
,
H.
,
Menniti
,
F.S.
,
Sobolevsky
,
A.I.
et al.
(
2021
)
Structure, Function, and Pharmacology of Glutamate Receptor Ion Channels
.
Pharmacol. Rev.
73
,
298
487
https://doi.org/10.1124/pharmrev.120.000131
51
Reiner
,
A.
and
Levitz
,
J
. (
2018
)
Glutamatergic Signaling in the Central Nervous System: Ionotropic and Metabotropic Receptors in Concert
.
Neuron
98
,
1080
1098
https://doi.org/10.1016/j.neuron.2018.05.018
52
Hansen
,
K.B.
,
Yi
,
F.
,
Perszyk
,
R.E.
,
Furukawa
,
H.
,
Wollmuth
,
L.P.
,
Gibb
,
A.J.
et al.
(
2018
)
Structure, function, and allosteric modulation of NMDA receptors
.
J. Gen. Physiol.
150
,
1081
1105
https://doi.org/10.1085/jgp.201812032
53
Paoletti
,
P.
,
Bellone
,
C.
and
Zhou
,
Q
. (
2013
)
NMDA receptor subunit diversity: impact on receptor properties, synaptic plasticity and disease
.
Nat. Rev. Neurosci.
14
,
383
400
https://doi.org/10.1038/nrn3504
54
Mota
,
S.I.
,
Ferreira
,
I.L.
,
Valero
,
J.
,
Ferreiro
,
E.
,
Carvalho
,
A.L.
,
Oliveira
,
C.R
, et al.
(
2014
)
Impaired Src signaling and post-synaptic actin polymerization in Alzheimer’s disease mice hippocampus—linking NMDA receptors and the reelin pathway
.
Exp. Neurol.
261
,
698
709
https://doi.org/10.1016/j.expneurol.2014.07.023
55
Fernandes
,
H.
and
Raymond
,
L
. (
2009
) NMDA receptors and Huntington’s disease.
In
Biology of the NMDA Receptor
,
CRC Press/Taylor & Francis
56
Molero
,
P.
,
Ramos-Quiroga
,
J.A.
,
Martin-Santos
,
R.
,
Calvo-Sánchez
,
E.
,
Gutiérrez-Rojas
,
L.
and
Meana
,
J.J
. (
2018
)
Antidepressant efficacy and tolerability of ketamine and esketamine: A critical review
.
CNS Drugs
32
,
411
420
https://doi.org/10.1007/s40263-018-0519-3
57
Mayer
,
M.L.
,
Westbrook
,
G.L.
and
Guthrie
,
P.B
. (
1984
)
Voltage-dependent block by Mg2+ of NMDA responses in spinal cord neurones
.
Nat. New Biol.
309
,
261
263
https://doi.org/10.1038/309261a0
58
Nowak
,
L.
,
Bregestovski
,
P.
,
Ascher
,
P.
,
Herbet
,
A.
and
Prochiantz
,
A
. (
1984
)
Magnesium gates glutamate-activated channels in mouse central neurones
.
Nat. New Biol.
307
,
462
465
https://doi.org/10.1038/307462a0
59
Lester
,
R.A.J.
,
Clements
,
J.D.
,
Westbrook
,
G.L.
and
Jahr
,
C.E
. (
1990
)
Channel kinetics determine the time course of NMDA receptor-mediated synaptic currents
.
Nat. New Biol.
346
,
565
567
https://doi.org/10.1038/346565a0
60
MacDermott
,
A.B.
,
Mayer
,
M.L.
,
Westbrook
,
G.L.
,
Smith
,
S.J.
and
Barker
,
J.L
. (
1986
)
NMDA-receptor activation increases cytoplasmic calcium concentration in cultured spinal cord neurones
.
Nat. New Biol.
321
,
519
522
https://doi.org/10.1038/321519a0
61
Mayer
,
M.L.
and
Westbrook
,
G.L
. (
1987
)
Permeation and block of N-methyl-D-aspartic acid receptor channels by divalent cations in mouse cultured central neurones
.
J. Physiol. (Lond.)
394
,
501
527
https://doi.org/10.1113/jphysiol.1987.sp016883
62
Nicoll
,
R.A.
and
Alger
,
B.E
. (
1981
)
Synaptic excitation may activate a calcium-dependent potassium conductance in hippocampal pyramidal cells
.
Science
212
,
957
959
https://doi.org/10.1126/science.6262912
63
Zorumski
,
C.F.
,
Thio
,
L.L.
,
Clark
,
G.D.
and
Clifford
,
D.B
. (
1989
)
Calcium influx through N-methyl-D-aspartate channels activates a potassium current in postnatal rat hippocampal neurons
.
Neurosci. Lett.
99
,
293
299
https://doi.org/10.1016/0304-3940(89)90462-x
64
Isaacson
,
J.S.
and
Murphy
,
G.J
. (
2001
)
Glutamate-mediated extrasynaptic inhibition: direct coupling of NMDA receptors to Ca(2+)-activated K+ channels
.
Neuron
31
,
1027
1034
https://doi.org/10.1016/s0896-6273(01)00428-7
65
Zhang
,
J.
,
Guan
,
X.
,
Li
,
Q.
,
Meredith
,
A.L.
,
Pan
,
H.L.
and
Yan
,
J
. (
2018
)
Glutamate-activated BK channel complexes formed with NMDA receptors
.
Proc. Natl. Acad. Sci. U.S.A.
115
,
E9006
E9014
https://doi.org/10.1073/pnas.1802567115
66
Gómez
,
R.
,
Maglio
,
L.E.
,
Gonzalez-Hernandez
,
A.J.
,
Rivero-Pérez
,
B.
,
Bartolomé-Martín
,
D.
and
Giraldez
,
T
. (
2021
)
NMDA receptor-BK channel coupling regulates synaptic plasticity in the barrel cortex
.
Proc. Natl. Acad. Sci. U.S.A.
118
, e2107026118 https://doi.org/10.1073/pnas.2107026118
67
Hayashi
,
Y.
,
Morinaga
,
S.
,
Liu
,
X.
,
Zhang
,
J.
,
Wu
,
Z.
,
Yokoyama
,
T.
et al.
(
2016
)
An EP2 Agonist Facilitates NMDA-Induced Outward Currents and Inhibits Dendritic Beading through Activation of BK Channels in Mouse Cortical Neurons
.
Mediators Inflamm.
2016
, 5079597 https://doi.org/10.1155/2016/5079597
68
Zhang
,
R.
,
Yamada
,
J.
,
Hayashi
,
Y.
,
Wu
,
Z.
,
Koyama
,
S.
and
Nakanishi
,
H
. (
2008
)
Inhibition of NMDA-induced outward currents by interleukin-1beta in hippocampal neurons
.
Biochem. Biophys. Res. Commun.
372
,
816
820
https://doi.org/10.1016/j.bbrc.2008.05.128
69
Zhang
,
R.
,
Sun
,
L.
,
Hayashi
,
Y.
,
Liu
,
X.
,
Koyama
,
S.
,
Wu
,
Z.
et al.
(
2010
)
Acute p38-mediated inhibition of NMDA-induced outward currents in hippocampal CA1 neurons by interleukin-1beta
.
Neurobiol. Dis.
38
,
68
77
https://doi.org/10.1016/j.nbd.2009.12.028
70
Ji
,
X.
,
Saha
,
S.
and
Martin
,
G.E
. (
2015
)
The origin of glutamatergic synaptic inputs controls synaptic plasticity and its modulation by alcohol in mice nucleus accumbens
.
Front. Synaptic Neurosci.
7
, 12 https://doi.org/10.3389/fnsyn.2015.00012
71
He
,
S.
,
Wang
,
Y.-X.
,
Petralia
,
R.S.
and
Brenowitz
,
S.D
. (
2014
)
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.
34
,
5261
5272
https://doi.org/10.1523/JNEUROSCI.3728-13.2014
72
Fan
,
F.
,
Chen
,
Y.
,
Chen
,
Z.
,
Guan
,
L.
,
Ye
,
Z.
,
Tang
,
Y.
et al.
(
2021
)
Blockade of BK channels attenuates chronic visceral hypersensitivity in an IBS-like rat model
.
Mol. Pain
17
, 17448069211040364 https://doi.org/10.1177/17448069211040364
73
Mitchell
,
D.E.
,
Miranda-Rottmann
,
S.
,
Blanchard
,
M.
and
Araya
,
R
. (
2023
)
Altered integration of excitatory inputs onto the basal dendrites of layer 5 pyramidal neurons in a mouse model of Fragile X syndrome
.
Proc. Natl. Acad. Sci. U.S.A.
120
, e2208963120 https://doi.org/10.1073/pnas.2208963120
74
Reyes-Carrión
,
A.
,
Bartolomé-Martín
,
D.
,
Rivero-Pérez
,
B.
and
Giraldez
,
T
. (
2023
)
NMDA RECEPTOR-BK CHANNEL FUNCTIONAL COUPLING IS ALTERED IN THE FRAGILE-X-SYNDROME MOUSE MODEL
.
IBRO Neurosci. Rep.
15
,
S331
S332
https://doi.org/10.1016/j.ibneur.2023.08.613
75
Horrigan
,
F.T.
and
Aldrich
,
R.W
. (
2002
)
Coupling between voltage sensor activation, Ca2+ binding and channel opening in large conductance (BK) potassium channels
.
J. Gen. Physiol.
120
,
267
305
https://doi.org/10.1085/jgp.20028605
76
Burnashev
,
N.
,
Zhou
,
Z.
,
Neher
,
E.
and
Sakmann
,
B
. (
1995
)
Fractional calcium currents through recombinant GluR channels of the NMDA, AMPA and kainate receptor subtypes
.
J. Physiol. (Lond.)
485
,
403
418
https://doi.org/10.1113/jphysiol.1995.sp020738
77
Martínez-Lázaro
,
R.
,
Minguez-Viñas
,
T.
,
Alvarez de la Rosa
,
D.
,
Bartolomé-Martín
,
D.
and
Giraldez
,
T
. (
2024
)
GRIN2B disease-associated mutations disrupt the function of BK channels and NMDA receptor signalling nanodomains
.
Physiol. (Bethesda).
https://doi.org/10.1101/2024.10.02.616357
78
Yaka
,
R.
,
Thornton
,
C.
,
Vagts
,
A.J.
,
Phamluong
,
K.
,
Bonci
,
A.
and
Ron
,
D
. (
2002
)
NMDA receptor function is regulated by the inhibitory scaffolding protein, RACK1
.
Proc. Natl. Acad. Sci. U.S.A.
99
,
5710
5715
https://doi.org/10.1073/pnas.062046299
79
Yang
,
J.-X.
,
Hua
,
L.
,
Li
,
Y.-Q.
,
Jiang
,
Y.-Y.
,
Han
,
D.
,
Liu
,
H.
et al.
(
2015
)
Caveolin-1 in the anterior cingulate cortex modulates chronic neuropathic pain via regulation of NMDA receptor 2B subunit
.
J. Neurosci.
35
,
36
52
https://doi.org/10.1523/JNEUROSCI.1161-14.2015
80
Isacson
,
C.K.
,
Lu
,
Q.
,
Karas
,
R.H.
and
Cox
,
D.H
. (
2007
)
RACK1 is a BKCa channel binding protein
.
Am. J. Physiol., Cell Physiol.
292
,
C1459
66
https://doi.org/10.1152/ajpcell.00322.2006
81
Wang
,
X.-L.
,
Ye
,
D.
,
Peterson
,
T.E.
,
Cao
,
S.
,
Shah
,
V.H.
,
Katusic
,
Z.S.
et al.
(
2005
)
Caveolae targeting and regulation of large conductance Ca(2+)-activated K+ channels in vascular endothelial cells
.
J. Biol. Chem.
280
,
11656
11664
https://doi.org/10.1074/jbc.M410987200
82
Roux
,
K.J.
,
Kim
,
D.I.
and
Burke
,
B
. (
2013
)
BioID: a screen for protein-protein interactions
.
Curr. Protoc. Protein Sci.
74
,
19
https://doi.org/10.1002/0471140864.ps1923s74
83
Fedele
,
L.
,
Newcombe
,
J.
,
Topf
,
M.
,
Gibb
,
A.
,
Harvey
,
R.J.
and
Smart
,
T.G
. (
2018
)
Disease-associated missense mutations in GluN2B subunit alter NMDA receptor ligand binding and ion channel properties
.
Nat. Commun.
9
,
957
https://doi.org/10.1038/s41467-018-02927-4
84
Lemke
,
J.R.
,
Hendrickx
,
R.
,
Geider
,
K.
,
Laube
,
B.
,
Schwake
,
M.
,
Harvey
,
R.J.
et al.
(
2014
)
GRIN2B mutations in West syndrome and intellectual disability with focal epilepsy
.
Ann. Neurol.
75
,
147
154
https://doi.org/10.1002/ana.24073
85
Vyklicky
,
V.
,
Krausova
,
B.
,
Cerny
,
J.
,
Ladislav
,
M.
,
Smejkalova
,
T.
,
Kysilov
,
B.
et al.
(
2018
)
Surface Expression, Function, and Pharmacology of Disease-Associated Mutations in the Membrane Domain of the Human GluN2B Subunit
.
Front. Mol. Neurosci.
11
, 110 https://doi.org/10.3389/fnmol.2018.00110
86
Ancatén-González
,
C.
,
Segura
,
I.
,
Alvarado-Sánchez
,
R.
,
Chávez
,
A.E.
and
Latorre
,
R
. (
2023
)
Ca2+- and Voltage-Activated K+ (BK) Channels in the Nervous System: One Gene, a Myriad of Physiological Functions
.
Int. J. Mol. Sci.
24
,
3407
https://doi.org/10.3390/ijms24043407
87
Hunt
,
D.L.
and
Castillo
,
P.E
. (
2012
)
Synaptic plasticity of NMDA receptors: mechanisms and functional implications
.
Curr. Opin. Neurobiol.
22
,
496
508
https://doi.org/10.1016/j.conb.2012.01.007
88
Guo
,
Y.
,
Liu
,
S.
,
Cui
,
G.-B.
,
Ma
,
L.
,
Feng
,
B.
,
Xing
,
J.
et al.
(
2012
)
Acute stress induces down-regulation of large-conductance Ca2+-activated potassium channels in the lateral amygdala
.
J. Physiol. (Lond.)
590
,
875
886
https://doi.org/10.1113/jphysiol.2011.223784
89
Pham
,
T.
,
Hussein
,
T.
,
Calis
,
D.
,
Bischof
,
H.
,
Skrabak
,
D.
,
Cruz Santos
,
M.
et al.
(
2023
)
BK channels sustain neuronal Ca2+ oscillations to support hippocampal long-term potentiation and memory formation
.
Cell. Mol. Life Sci.
80
,
369
https://doi.org/10.1007/s00018-023-05016-y
90
Park
,
S.M.
,
Roache
,
C.E.
,
Iffland
,
P.H.
2nd
,
Moldenhauer
,
H.J.
,
Matychak
,
K.K.
,
Plante
,
A.E.
et al.
(
2022
)
BK channel properties correlate with neurobehavioral severity in three KCNMA1-linked channelopathy mouse models
.
Elife
11
, e77953 https://doi.org/10.7554/eLife.77953
91
Groc
,
L.
,
Heine
,
M.
,
Cousins
,
S.L.
,
Stephenson
,
F.A.
,
Lounis
,
B.
,
Cognet
,
L.
et al.
(
2006
)
NMDA receptor surface mobility depends on NR2A-2B subunits
.
Proc. Natl. Acad. Sci. U.S.A.
103
,
18769
18774
https://doi.org/10.1073/pnas.0605238103
92
Song
,
B.
and
Marvizón
,
J.C.G
. (
2005
)
N-methyl-D-aspartate receptors and large conductance calcium-sensitive potassium channels inhibit the release of opioid peptides that induce mu-opioid receptor internalization in the rat spinal cord
.
Neuroscience
136
,
549
562
https://doi.org/10.1016/j.neuroscience.2005.08.032
This is an open access article published by Portland Press Limited on behalf of the Biochemical Society and distributed under the Creative Commons Attribution License 4.0 (CC BY).