Next Article in Journal
Cold Atmospheric Plasma Does Not Affect Stellate Cells Phenotype in Pancreatic Cancer Tissue in Ovo
Next Article in Special Issue
Neuronal Roles of the Multifunctional Protein Dipeptidyl Peptidase-like 6 (DPP6)
Previous Article in Journal
Takotsubo Syndrome: Translational Implications and Pathomechanisms
Previous Article in Special Issue
Inhibition of the Akt/PKB Kinase Increases Nav1.6-Mediated Currents and Neuronal Excitability in CA1 Hippocampal Pyramidal Neurons
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Ion Channel Partnerships: Odd and Not-So-Odd Couples Controlling Neuronal Ion Channel Function

by
Nicholas C. Vierra
and
James S. Trimmer
*
Department of Physiology and Membrane Biology, University of California Davis School of Medicine, Davis, CA 95616, USA
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(4), 1953; https://doi.org/10.3390/ijms23041953
Submission received: 25 January 2022 / Revised: 5 February 2022 / Accepted: 8 February 2022 / Published: 10 February 2022
(This article belongs to the Special Issue Neuronal Voltage-Gated Channel Regulation)

Abstract

:
The concerted function of the large number of ion channels expressed in excitable cells, including brain neurons, shapes diverse signaling events by controlling the electrical properties of membranes. It has long been recognized that specific groups of ion channels are functionally coupled in mediating ionic fluxes that impact membrane potential, and that these changes in membrane potential impact ion channel gating. Recent studies have identified distinct sets of ion channels that can also physically and functionally associate to regulate the function of either ion channel partner beyond that afforded by changes in membrane potential alone. Here, we review canonical examples of such ion channel partnerships, in which a Ca2+ channel is partnered with a Ca2+-activated K+ channel to provide a dedicated route for efficient coupling of Ca2+ influx to K+ channel activation. We also highlight examples of non-canonical ion channel partnerships between Ca2+ channels and voltage-gated K+ channels that are not intrinsically Ca2+ sensitive, but whose partnership nonetheless yields enhanced regulation of one or the other ion channel partner. We also discuss how these ion channel partnerships can be shaped by the subcellular compartments in which they are found and provide perspectives on how recent advances in techniques to identify proteins in close proximity to one another in native cells may lead to an expanded knowledge of other ion channel partnerships.

1. Introduction

Neuronal voltage-gated ion channels interact with a constellation of proteins that impact their activity, subcellular localization, and associated signaling pathways. Diversity in the gating properties and physiological roles of many voltage-gated ion channels is generated by interactions of the channel’s pore-forming α subunits with auxiliary subunits that control channel activity and membrane trafficking. Voltage-gated ion channels also physically associate with numerous cytoskeletal, anchoring, and scaffolding proteins that play diverse roles in impacting their expression, localization, and function, and that are essential in organizing discrete ion channel signaling systems. The assembly of distinct ion channel signaling complexes and their subcellular distribution are fundamental determinants of neuronal function [1]. In addition to interactions with diverse auxiliary subunits and non-ion channel proteins, some voltage-gated ion channels rely on the partnership with other ion channels to carry out their physiological functions. These ion channel partnerships work to promote efficient coupling of the distinct functions of the partner ion channels to one another and expand the diversity of the physiological events that they underlie.
The indirect coordination of voltage-gated ion channel function through the effects of their ionic currents on membrane potential is well-appreciated. However, close physical association of specific ion channel pairs can facilitate more efficient functional coupling, resulting in more rapid and reliable activation and modulation of their respective functions. Many examples of ion channel partnerships critical for regulating electrical activity are those that work to control and/or quickly respond to elevations in intracellular Ca2+. Because Ca2+ influx is coupled to diverse signaling pathways that can have profound and sometimes opposing effects on neuronal function [2], and that aberrantly elevated intracellular Ca2+ is toxic to cells [3], the sites of Ca2+ influx and its intracellular concentration are tightly regulated [4]. Precise spatiotemporal control of Ca2+ influx into compartmentalized signaling domains permits rapid and selective activation of Ca2+-dependent effectors localized in specialized nanodomains no more than a few hundred nanometers from the Ca2+ source [5,6]. The partnership of Ca2+-permeable ion channels with Ca2+-dependent effectors enables spatially and temporally precise activation of Ca2+-dependent processes, including regulation of ion channel function. In neurons, elevations in intracellular Ca2+ are generated primarily by the opening of ligand-gated, store-operated, and voltage-gated Ca2+ channels (VGCCs) in the plasma membrane (PM), and organellar ion channels localized predominantly in the endoplasmic reticulum (ER) [7].
The molecular composition and subcellular distribution of VGCC-associated signaling complexes is a key mechanism for regulating VGCC-mediated Ca2+ signals in neurons [1,8,9]. Early biophysical characterization of VGCCs before their molecular identities were known also led to an alphabetical classification based on their current and gating properties (Table 1). VGCCs are defined at the molecular level by their principal or α1 subunit, which contains the pore-forming and voltage-sensing domains [8]. The 10 genes encoding VGCC α1 subunits in mammals are divided into three phylogenetic subgroups, including the high-voltage activated Cav1 and Cav2 channels, and the low-voltage activated Cav3 channels [8]. All VGCC α1 subunit genes give rise to splice variants that can alter VGCC properties. The α1 subunit of neuronal Cav1 and Cav2 channels assembles with non-conducting auxiliary α2δ and β subunits that influence VGCC membrane trafficking and gating [8].
While VGCCs influence neuronal electrical activity directly through effects on membrane potential via the Ca2+ currents they pass, they also control neuronal function by coupling electrical activity to intracellular Ca2+ signals that trigger diverse physiological responses. Each of the VGCC subgroups has a characteristic predominant subcellular localization in neurons that defines their distinct contributions to neuronal function [10]. For example, L-type Cav1 channels (LTCCs) are prominently found on the cell body (or soma) and proximal dendrites where they are positioned to mediate Ca2+ signals that can rapidly induce changes in gene expression [11,12,13], and also on distal dendrites at or near dendritic spines where they can modulate synaptic function [14,15]. Cav2 channels are a critical component of the synaptic vesicle release machinery and are enriched in presynaptic terminals [16,17]. The physiological properties and functions of Cav3 channels are not as well-studied as those of Cav1 and Cav2 channels, but important roles for Cav3 channels have been identified in regulating electrical excitability in many different types of brain neurons [18]. De novo pathogenic variants in all classes of VGCCs have been linked to debilitating neurological disorders, highlighting the key contributions of these channels to normal brain function [19]. The interactions of VGCC with cytoskeletal, anchoring, and exocytotic machinery proteins are intimately involved in organizing discrete Ca2+ signaling systems in different neuronal compartments. For example, the interaction of VGCCs with the synaptic vesicle exocytosis machinery in axon terminals is a prominent and well-explored example of an ion channel–protein partnership that is crucial to brain function [20]. In addition to these interactions, the partnership of VGCCs with numerous other ion channels provides a key mechanism to functionally organize VGCC-dependent Ca2+ signaling processes.

2. Canonical Ion Channel Partnerships

Functional ion channel partnerships established by a close spatial association of ion channel pairs or higher-order oligomers have been reported for many voltage-gated ion channels. Homomeric ion channel complexes are described for VGCCs, voltage-gated Na+ channels, and voltage-gated K+ channels (for a detailed review, see [21]). In many such cases, channel oligomerization serves a key role in permitting physical interactions between channels to modulate gating properties, tuning electrical activity. For example, the clustering of Cav1.2 channels promotes physical interactions between adjacent Cav1.2 monomers, enabling their functional coupling and promoting cooperative gating of Cav1.2 channels [22,23]. Below, we describe ion channel partnerships formed between distinct ion channels and detail the key physiological roles linked to these associations.

2.1. PM-ER Ion Channel Partnerships

Neuronal Ca2+ signals initiated by Ca2+ influx through PM VGCCs can be accompanied by ER Ca2+ release, amplifying the Ca2+ signal beyond that produced by opening of the PM Ca2+ channel alone. The storage and release of Ca2+ is a fundamental function of the ER, and some regions of the neuronal ER are formed into specialized structures known as subsurface cisternae. Subsurface cisternae come into close contact with the PM (within 20 nm) [24], are prominent in the soma of many brain neurons [25], and in some neurons are enriched with ER Ca2+ release channels [26,27]. Ca2+ signals arising from Ca2+-permeable channels found on the neuronal ER are generated primarily by inositol 1,4,5-trisphosphate receptors (IP3Rs) and ryanodine receptors (RyRs). IP3R-produced Ca2+ signals result from the production of IP3 generated by activation of phospholipase C. RyRs are gated by elevations in cytosolic Ca2+, releasing ER Ca2+ in a process known as Ca2+-induced Ca2+ release (CICR). RyR-dependent CICR can be triggered by Ca2+ entering the cytoplasm through PM-localized Ca2+-permeable channels, including VGCCs, especially those localized at PM sites near ER-localized RyRs. RyR-generated Ca2+ signals couple to distinct physiological processes in neurons, including regulation of synaptic signaling [28], regulation of gene expression [29], morphological plasticity of dendritic spines in hippocampal pyramidal neurons [30], and the regulation of electrical excitability via activation of PM-localized Ca2+-activated K+ channels [31,32,33]. While VGCC-dependent activation of neuronal RyRs has been reported for Cav3 channels [34,35], most reports in neurons describing the functional coupling of PM VGCCs with RyRs involve Cav1 and Cav2 channels. We note that, in most cases, the specific isoforms of RyRs (RyRs 1-3) participating in these partnerships have not been defined.

2.2. Cav1-RyR

In cardiomyocytes, where the partnership of Cav1.2 channels with RyR2 is well-explored, the opening of Cav1.2 channels localized at junctional dyads, a form of ER–PM junctions, triggers robust CICR [36]. Although the physiological responses to Cav1-RyR-mediated Ca2+ signals differ between cardiomyocytes and neurons, the key molecular attributes of Cav1-RyR coupling enabling CICR appear to be shared between neurons and non-neuronal cells. Like cardiomyocytes, membrane potential depolarization increases the frequency of spontaneous Ca2+ sparks, or ER Ca2+ release events [37], in the soma and proximal dendrites of hippocampal pyramidal neurons [38,39]. Pharmacological blockade of either Cav1 channels or RyRs block these Ca2+ sparks, indicating that Ca2+ influx through both channel types is required for neuronal Ca2+ spark generation [38,39,40].
Although the specific physiological functions of spontaneous Ca2+ sparks produced by the neuronal Cav1–RyR partnership are unclear, the functional coupling of these channels requires their juxtaposition at ER–PM junctions. In cardiomyocytes, the distance from the mouth of the Cav1.2 channel and RyRs is around 10 nm [41]. A considerable amount of evidence suggests a similar spatial arrangement for functionally coupled Cav1 channels and RyRs in mammalian brain neurons. Early measurements of VGCC-dependent CICR in the soma of cultured hippocampal neurons using confocal microscopy revealed that this form of Ca2+ signaling occurs within 3 µm of the PM [42], and similar observations have been made in cartwheel cells of the cochlear nucleus [33]. Cav1 channel-dependent CICR can be inhibited by ER-resident STIM1 molecules [43,44], indicating that Cav1 channel signaling complexes come into close apposition with the ER [45]. Finally, Vierra and colleagues [40] demonstrated that Cav1 channels co-immunopurify with the voltage-gated K+ channel Kv2.1 and RyR-containing protein complexes isolated from chemically crosslinked mouse brain homogenates (Figure 1). These Kv2.1-organized signaling complexes at neuronal ER–PM junctions are responsible for localized spontaneous Ca2+ sparks in the soma of hippocampal CA1 pyramidal neurons [40].

2.3. VGCC–RyR–KCa

RyR-mediated intracellular Ca2+ release couples to a wide range of cellular responses, including activation of Ca2+-activated K+ (KCa) channels. There exist three families of KCa channels (large conductance KCa1, small conductance KCa2, and intermediate conductance KCa3) whose members play an essential role modulating neuronal action potential firing patterns [46]. Many examples of RyR–KCa partnerships in neurons also include VGCCs as a third partner. By opening upon membrane potential depolarization VGCCs provide an action potential stimulus-dependent trigger for CICR leading to activation of KCa channels, which then contribute to repolarizing the membrane potential to affect subsequent action potential firing.
Members of the small-conductance Ca2+-activated K+ (SK) channel family are widely expressed in brain neurons where they are important for regulation intrinsic excitability and firing patterns. SK channels are indirectly activated by sub-micromolar elevations in intracellular Ca2+ through a requisite physical interaction with calmodulin, which functions as the Ca2+ sensor for SK channels (for a detailed review, see [47]). SK channels display heterogeneous and distance-dependent functional coupling to various Ca2+-permeable channels in neurons, including VGCCs and RyRs. For example, in cerebellar Purkinje neurons, SK channels are activated by Cav2.1-triggered CICR through RyRs at ER–PM junctions [48]. Double knockout of the junctophilin-3 and -4 proteins, which stabilize ER–PM contacts in these cells, results in impaired electrical activity and motor dysfunction [48].
In contrast to SK channels, the large-conductance Ca2+-activated K+ channel KCa1.1 (i.e., the BK channel) has a lower affinity for Ca2+ and must be localized much closer to the Ca2+ source than SK channels to ensure reliable Ca2+-dependent activation [49]. Consequently, gating of BK channels is highly dependent on the physical association of BK channels with the Ca2+-permeable channels that provide their Ca2+ source. Kaufman and colleagues used immunogold electron microscopy to demonstrate that somatic BK channels in cerebellar Purkinje neurons cluster over subsurface cisternae ER–PM junctions enriched in IP3Rs [26]. They subsequently showed that, in many types of brain neurons, somatic BK channels exist in PM clusters [50], although the relationship of these clustered BK channels to ER–PM junctions and other intracellular Ca2+ sources was not determined. RyR-dependent gating of BK channels serves to promote membrane hyperpolarization, dampening electrical activity and/or shaping action potential firing patterns. As a prominent recent example, in cartwheel interneurons of the dorsal cochlear nucleus, where action potential bursts are a key signaling mechanism, Irie and Trussell [33] found that RyR-mediated Ca2+ signals triggered by closely associated Cav2.1 channels activate somatic BK channels (Figure 2B). The Cav2.1–RyR–BK triad provides these neurons with a mechanism to ensure rapid and reliable gating of BK channels to control action potential firing patterns [33]. Indriati and colleagues [51] also reported the colocalization of Cav2.1 and BK channels at somatic subsurface cisternae in cerebellar Purkinje cells, suggesting that the Cav2.1–RyR–BK triad may be present across distinct types of neurons to regulate electrical activity.

2.4. VGCC–BK

A close spatial association of BK channels with VGCCs was first inferred from biophysical evidence reported over 30 years ago. Experiments performed in cells loaded with Ca2+ chelators, which buffer elevations in intracellular Ca2+ and restrict the spatial extent of its action from its source, revealed that some BK channels are tightly associated with intracellular Ca2+ sources [52,53]. In addition, the sensitivity of BK currents to pharmacological blockers of different VGCCs can vary according to the subcellular compartment in which they are recorded, indicating the existence of distinct compartment-specific VGCC–BK channel partnerships [54]. These observations led to an effort to identify the specific molecular mechanisms that link BK channels to specific Ca2+ sources in neurons.
The functional coupling of VGCCs with BK channels can be reconstituted in heterologous expression systems, suggesting a direct molecular interaction of BK channels with Cav1 and Cav2 VGCCs. In support of this model, mass spectrometry-based analyses of protein complexes immunopurified from mouse and rat brain using BK- and Cav1.2- [55] or Cav2- and Cavβ- [56] specific antibodies revealed that BK channels co-purify with Cav1.2 and Cav2 α1 subunits, VGCC auxiliary subunits, as well as numerous other Ca2+-dependent signaling molecules. The results of these proteomics experiments combined with parallel coexpression studies in heterologous cells provided a detailed molecular mechanism explaining how Cav2 channels functionally couple to BK channels in presynaptic terminals [57] where BK currents limit neurotransmitter release (Figure 2C). These studies also revealed that the close spatial proximity of these ion channel partners resulted in a dedicated nanodomain for the Ca2+ entering through the VGCCs to trigger activation of the associated BK channels, as shown by the distinct sensitivity of BK channel activation to BAPTA but not EGTA Ca2+ chelators [55].
Vivas and colleagues used super-resolution microscopy and immunolabeling to reveal a molecular interaction between Cav1.3 and BK channels in both heterologous cells expressing exogenous channels, and on the somata of cultured neurons, identifying a mechanism whereby BK channels can quickly influence action potential firing through their activation by Cav1.3-generated Ca2+ signals (Figure 2D) [58]. In hippocampal CA1 pyramidal neurons, Gutzmann and colleagues found that somatodendritic BK channels physically and functionally couple with Cav2.3 channels to control intrinsic excitability (Figure 2D) [59]. While loss of depolarizing Ca2+ currents through Cav2.3 would be predicted to dampen electrical excitability, loss of Cav2.3 channels instead results in increased excitability. This was found to result from reduced BK and SK channel K+ currents whose activity depends on Cav2.3-mediated Ca2+ influx. These results highlight the powerful impact ion channel partnerships can have on the physiological outputs of individual ion channels, and that these partnerships can lead to functional consequences that are not always predictable based on the physiological role of either individual partner.
Interestingly, in some neurons, BK channels can dynamically couple to distinct Ca2+ sources. Whitt and colleagues identified diurnal modulation of BK channels in neurons of the suprachiasmatic nucleus, functionally coupling BK channels to Cav1 channels during the day while favoring their activation by RyRs at night when Cav1 currents are attenuated [60]. These results suggest that regulation of the Ca2+ source may be a key mechanism whereby neurons tune the response properties of BK channels engaged in such partnerships. It may be that some VGCC–BK channel partnerships are modulated by additional effectors that transiently affect their functional properties or spatial distribution, for example by dynamically regulating their distance from intracellular Ca2+ sources through changes in their PM clustering state.

2.5. VGCC–IK

The functional coupling of intermediate-conductance Ca2+-activated K+ (IK) channels to VGCC-dependent Ca2+ entry has also been reported in brain neurons. Like SK channels, IK channels are indirectly gated by intracellular Ca2+ via an interaction with calmodulin [61]. They are distinguished from SK channels by their larger K+ conductance and distinct pharmacological sensitivities. IK channels were shown to contribute to the slow afterhyperpolarization, a Ca2+-dependent K+ current that contributes to the regulation of intrinsic excitability and firing patterns [62], although there has been debate over the molecular identity of the channel(s) that generate the slow afterhyperpolarization due to its unique biophysical and pharmacological properties [63]. Sahu and colleagues reported the preferential functional coupling of IK channels with Cav1.3 channels in hippocampal CA1 pyramidal neurons [64]. Spatial and functional coupling of Cav1.3 with IK channels could be reconstituted in heterologous cells when the scaffolding protein densin and Ca2+/calmodulin-dependent kinase II (CaMKII) were also present. This same group found that Cav1.3, IK, and RyR2 channels colocalized at somatic ER–PM junctions stabilized by junctophilin-3 and -4 proteins in hippocampal CA1 pyramidal neurons (Figure 3) [65]. Knockdown of these junctophilin proteins impaired generation of the slow afterhyperpolarization, indicating that in CA1 pyramidal neurons, junctophilin-3 and -4 containing ER–PM junctions serve an important role promoting this particular ion channel partnership [65].

3. Non-Canonical Ion Channel Partnerships

Over the last decade, several studies have revealed the existence of non-canonical ion channel partnerships. In these cases, two ion channels that do not have an obvious direct relationship in terms of their respective functions form physical and functional associations. As in many of the canonical partnerships summarized above, these non-canonical examples also involve VGCC channels. However, in these cases, the pairings are with voltage-gated Kv channels that are not intrinsically Ca2+ sensitive. However, these physical and functional pairings result in regulation of one or the other channel partner to underlie important physiological events.

3.1. VGCC–Kv Channel Partnerships

Functional coupling of VGCCs and voltage-gated K+ (Kv) channels as mediated by their respective impacts on membrane potential is fundamental to signaling in brain neurons. It has long been appreciated that the regulated cellular coexpression and subcellular colocalization of specific combinations of VGCCs and Kv channels underlie diverse neuronal signaling events. However, there is a growing understanding that physical and functional associations between VGCC and Kv channel partners can serve important roles in neuronal signaling beyond their canonical functions as defined by the respective channels’ separate effects on membrane potential.
Kv channels are highly diverse and like VGCCs, Kv channels are primarily defined by their α subunit, four of which coassemble to form a functional channel. Depending on the Kv channel, the core α subunit tetramer can assemble with Kv subclass-specific cytoplasmic and transmembrane auxiliary subunits, which can dramatically impact Kv channel expression, localization, and function.

3.1.1. Cav3–Kv4

The first example defined in which there exists a physical and functional association between a VGCC and a Kv channel was that between low voltage-activated Cav3 T-type Ca2+ channels and the Kv4 K+ channels that mediate a major component of transient potassium currents (IA) in many brain neurons. In cerebellar stellate cells, this non-canonical pairing of two voltage-gated ion channels leads to enhancement of IA in the subthreshold voltage range by T-type Ca2+ channel-mediated Ca2+ influx [66]. The authors found that Ca2+ entry via Cav3 channels leads to a shift in the voltage-dependence of inactivation of IA towards more positive potentials without affecting activation. This resulted in enhanced IA due to the reduced steady state inactivation of IA which, combined with the lack of prominent effects on activation, results in an increase in the IA window current. Biochemical analyses showed that Cav3.2- and Cav3.3-containing T-type Ca2+ channel protein complexes immunopurified from rat brain contained Kv4.2 α subunit and KChIP3 auxiliary subunit polypeptides, known to be component subunits of a prominent population of transient potassium channels found in brain neurons [67]. The KChIP auxiliary subunits of Kv4 channels are members of the EF-hand domain-containing neuronal Ca2+ sensor protein family [68], although the role of KChIPs acting as Ca2+ sensors to confer physiological Ca2+-mediated modulation to native Kv4 channel function in neurons had not been established. The authors were able to reconstitute the T-type Ca2+ channel-mediated modulation of Kv4-based IA in experiments in transfected heterologous cells coexpressing these partner ion channels. They then used this system to show that KChIP3 coexpression was required for such modulation. To assess the contribution of KChIP subunits to the T-type Ca2+ channel-mediated enhancement of IA in native cells, the authors dialyzed anti-KChIP monoclonal antibodies [69] into stellate cells via a patch pipet. Dialysis of anti-KChIP3 antibodies in particular eliminated the effects of T-type channel-mediated Ca2+ entry on inactivation of native neuronal IA, mimicking the effects of blocking T-type channels.
This unique coupling of these two voltage-sensitive ion channels was subsequently shown by the same group to underlie the gain in stellate cell inhibitory output in response to reduced extracellular [Ca2+], such as what occurs during repetitive climbing fiber stimulation that yields complex spiking in cerebellar Purkinje cells [70]. The authors found that experimentally reducing extracellular [Ca2+] led to enhanced steady-state inactivation of the Kv4-based stellate cell IA, similar to what is seen upon inhibiting Ca2+ entry through Cav3 channels. Similar effects of reducing extracellular [Ca2+] in the recording solution were also seen in heterologous cells expressing components of the Cav3–Kv4 complex, with the functional coupling between the VGCC and Kv channel again dependent on coexpression of the KChIP3 auxiliary subunit of the Kv4 channels. Repetitive climbing fiber stimulation leading to complex spike discharge in cerebellar Purkinje cells mimicked the effects of experimentally reducing extracellular [Ca2+] on stellate cell Cav3-mediated Ca2+ influx and reduced the amplitude of the Kv4-based IA. Detailed pharmacological analyses, including the use of the anti-KChIP3 monoclonal antibody, defined the Cav3–Kv4 complex as being the mediator of these responses. The overall effect of reducing extracellular [Ca2+] is to reduce the enhancement of stellate cell IA mediated by the Cav3–Kv4 complex (Figure 4A), resulting in increased stellate cell firing rate to sustain inhibitory synaptic input onto Purkinje cells and maintaining cerebellar circuit function during high-frequency stimulation.

3.1.2. Cav2.3–Kv4.2

There are numerous forms of excitatory synaptic potentiation that underlie circuit and behavioral plasticity. Excitatory synaptic potentiation has been well studied in the hippocampal Schaeffer collateral pathway through which CA3 pyramidal neurons provide excitatory input onto CA1 pyramidal cells. One form of experimental postsynaptic excitatory synaptic potentiation in hippocampal CA1 pyramidal neurons is mediated by the partnership of the high voltage-activated Cav2.3 R-type Ca2+ channel with the Kv4.2 Kv channel (Figure 2C) [71]. Pharmacological blockade of Cav2.3-mediated Ca2+ influx with the tarantula neurotoxin SNX-482 (SNX) led to dynamic boosting of excitatory postsynaptic potentials (EPSPs) in hippocampal CA1 pyramidal neurons [71]. While SNX can also directly inhibit Kv4 channel gating [72], this EPSP boosting was sensitive to the fast Ca2+ chelator BAPTA, suggesting that Ca2+ entry through SNX-sensitive Cav2.3 channels is coupled to a downstream mediator located in close proximity. This Cav2.3-based pathway was unique from a previously defined signaling pathway whereby Ca2+ entry through NMDA receptors activated apamin-sensitive small conductance Ca2+-activated K+ or SK channels (Figure 2C) to yield dampening of EPSPs [73]. The SNX-induced EPSP boosting was also inhibited by 4-AP used at concentrations where it is a fairly selective blocker of Kv4-containing K+ channels. Exogenous expression of a dominant negative isoform of Kv4.2 and dialysis of anti-KChIP monoclonal antibodies through a patch pipet also abolished the SNX-induced boosting of EPSPs. This supported a role for KChIP-containing Kv4.2 channels in coupling the inhibition of Ca2+ entry through SNX-sensitive Cav2.3 R-type Ca2+ channel activity to EPSP boosting. The authors proposed a model that a Cav2.3–Kv4.2 complex in the dendrites of CA1 pyramidal neurons (Figure 4B) functions in a manner analogous to the Cav3–Kv4 complex in cerebellar stellate cells. In this model, Ca2+ entry through Cav2.3 channels enhances Kv4-based IA availability, which then serves to dampen EPSP amplitude.
It should be noted that the authors also defined a distinct mechanism for dynamic regulation of CA1 pyramidal cell EPSP amplitude through the more canonical functional coupling of a Ca2+-permeable ion channel, the NMDA-type ionotropic glutamate receptor, to a Ca2+-activated SK channel (Figure 2C). It is intriguing that immunogold electron microscopy studies of the localization of these ion channels in dendritic spines of CA1 neurons demonstrate distinct pairwise spatial coupling of the two distinct ion channel pairs that underlie the canonical and non-canonical partnerships. Labeling for both SK channels and NMDA receptors is found within the postsynaptic density [74], whereas that for both Cav2.3 [75] and Kv4.2 [76] is found at sites within spines outside of the postsynaptic density itself (Figure 2C). Subsequent studies employing Cav2.3 knockout mice showed that the effects of SNX to boost the amplitude of CA1 pyramidal cell EPSPs were mediated through Cav2.3 channels [77,78] and not through the reported [72] off-target effects of SNX on the Kv channels mediating IA.
While the work of Wang and colleagues indicated that Kv4.2 channels and Cav2.3 channels are in close proximity, Murphy and colleagues recently demonstrated that these two channels physically associate in a protein complex. They identified Cav2.3 channels as molecular partners of Kv4.2 channels by using tandem affinity purification combined with mass spectrometry to isolate and analyze Kv4.2-containing protein complexes from hippocampal neurons [79]. Native Kv4.2 channels co-immunoprecipitated with Cav2.3 in wild-type neurons but not Cav2.3 KO neurons, and imaging experiments using fluorescently tagged Kv4.2 and Cav2.3 channels in hippocampal neurons confirmed that Kv4.2 and Cav2.3 channels form stable complexes in dendrites and at excitatory synapses. Moreover, they determined that Cav2.3 and Kv4.2 likely engage in a direct interaction at a 1:1 channel stoichiometry, enabling Ca2+ influx through Cav2.3 to gate the associated Kv4.2–KChIP complex. Acute blockade of Cav2.3 channels using Ni2+ reduced IA and decreased surface localization of Kv4.2 channels, suggesting that the Cav2.3–Kv4.2 partnership provides an activity-dependent mechanism to dynamically regulate dendritic excitability. Finally, they demonstrated that the dendritic IA gradient seen in CA1 neurons [80] was altered in neurons from mice lacking Cav2.3 channels, disrupting IA modulation of synaptic currents. These results indicate that the partnership of Kv4.2 and Cav2.3 channels is a key mechanism for regulating synaptic inputs in hippocampal neurons.
The Cav2.3–Kv4.2 partnership seen in glutamatergic CA1 hippocampal pyramidal neurons is not detectable in glutamatergic cerebellar granule cells, which instead have the Cav3–Kv4 partnership previously defined in GABAergic cerebellar stellate cells [81]. Interestingly, there was an anterior–posterior gradient of the amplitude of Cav3 Ca2+ current as measured in granule cells in different lobules of the cerebellum, with higher Cav3 Ca2+ current amplitudes posterior versus anterior. This resulted in a more prominent functional role for the Cav3–Kv4 partnership in posterior cerebellum. It is intriguing that the Kv4.2 and Kv4.3 α subunits also exhibit an anterior–posterior expression gradient at the mRNA [82], protein [83] and functional [84] level, with Kv4.3 predominating in posterior lobes and Kv4.2 in anterior lobes. This provides a compelling example whereby the patterns of cellular expression of the ion channel partners across the cerebellum can profoundly impact firing in these different populations of granule cells in response to their distinct forms of synaptic input.

3.2. Ion Channel Partnerships Involving Non-Conducting Functions of Ion Channels

Voltage-gated ion channel partnerships that utilize the non-conducting functions of a constituent channel such as its scaffolding and/or subcellular targeting properties are not as well-studied as partnerships dependent upon regulation conferred by juxtaposed ionic currents. A prominent example of a partnership that relies upon a non-conducting function of an ion channel is present in skeletal muscle, where the physical interaction of PM Cav1.1 LTCCs with RyR1 channels in the sarcoplasmic reticulum is essential for excitation-contraction coupling. Unlike the coupling of Cav1.2 channels and RyR2 channels in cardiomyocytes, Ca2+ conductance through Cav1.1 is not required to trigger RyR-mediated CICR in skeletal muscle [85]. Instead, Cav1.1 channels function as a voltage sensor for RyRs, allowing RyRs to open in response to voltage-induced conformational changes in Cav1.1. Thus, the Cav1.1-RyR1 pairing represents an example of an ion channel partnership that relies on a non-conducting role of one of the channels to impact the function of another through a physical association. In neurons, the partnership of the voltage-gated K+ channel Kv2.1 with Cav1.2 and RyRs has emerged as an ion channel partnership dependent upon the non-conducting roles of Kv2.1.

Cav1.2–Kv2.1

Kv2.1 is expressed at high levels in brain neurons where it is present in large clusters that are prominent on the soma and proximal dendrites, and on the axon initial segment [67]. The sites where Kv2.1 forms clusters in both neurons and in heterologous cells are ER–PM junctions [27,86,87,88,89]. Fox, Tamkun and colleagues showed that not only is Kv2.1 localized to ER–PM junctions, it plays an active role in organizing these membrane contact sites [90]. Subsequent studies have revealed that PM Kv2.1 channels interact with ER VAP proteins to form and maintain these membrane contact sites (Figure 1) [91,92], a nonconducting function of Kv2.1 and its paralog Kv2.2 [93], and mediated by the “PRC domain” [94] present on the cytoplasmic C-terminus of both of these Kv2 α subunits.
Endogenous Kv2.1 channels in cultured hippocampal neurons and in brain neurons are found in close spatial association with clusters of Cav1.2 VGCCs [40]. Studies employing coexpression in heterologous cells revealed that coexpression with a nonconducting mutant of Kv2.1 enhances clustering of Cav1.2 at Kv2.1-containing ER–PM junctions. This results in an increase in LTCC activity and enhanced CICR. The former is a consequence of enhanced cooperative gating of LTCCs upon their coclustering [22], and the latter of the increased spatial and function association of the VGCCs with ER RyRs upon their coclustering with Kv2.1 at ER–PM junctions. Neurons from Kv2.1 KO mice have correspondingly reduced Cav1.2 clustering, and reduced spatial and functional association of LTCCs and RyRs [40].
Studies in vascular smooth muscle cells further supported a role for Kv2.1 in regulating the clustering and activity of LTCCs [95]. O’Dwyer, Santana and colleagues found that myocytes from female mice have larger Cav1.2 clusters, higher intracellular [Ca2+], and increased myogenic tone compared to those from male myocytes. These differences paralleled the level of expression of Kv2.1 in female (high) versus male (low) myocytes. Moreover, these sex-specific differences in the clustering and activity of LTCCs were eliminated in myocytes from Kv2.1 KO mice. This study supports a role for Kv2.1-mediated clustering of LTCCs as an important determinant of Ca2+ signaling in diverse excitable cell types.
These studies used differences in Kv2.1 expression to gain insights into its role in regulating LTCCs and their coupling to intracellular Ca2+ signaling. While not directly addressed experimentally, it is likely that given the role of Kv2.1 in forming and maintaining ER–PM junctions, the overall organization of ER–PM junctions was also impacted in these experimental systems exhibiting differences in Kv2.1 expression. To try to separate the broader role of Kv2.1 in forming ER–PM junctions from its ability to recruit LTCCs to these sites, Vierra and colleagues set out to develop tools to selectively disrupt the latter but not the former [96]. They identified a motif, the “Ca2+ channel association domain” or CCAD on the proximal C-terminus of Kv2.1, that was necessary and sufficient for clustering LTCCs at ER–PM junctions. Kv2.1 mutants lacking or mutated within the CCAD were still fully competent to form clusters at and organize ER–PM junctions but they failed to recruit LTCCs to these sites. These characteristics are distinct from mutants in the downstream PRC domain which failed to form ER–PM junctions [91,92,93]. The authors also developed a cell-penetrating interfering peptide based on the CCAD sequence that also eliminated LTCC clustering at Kv2.1-containing ER–PM junctions without impacting these membrane contact sites themselves [96]. The authors used proximity ligation assays [97], which detect protein pairs within 40 nm of one another [98], employing antibodies against Kv2.1 and the Cav1.2 LTCC subunit, to define the homotypic and heterotypic coclustering of Kv2.1 and Cav1.2. They found that these CCAD-based interventions eliminated the spatial association of Cav1.2 with Kv2.1 as well as the Kv2.1-mediated enhanced coclustering of Cav1.2 channels with one another. In addition, these CCAD-based interventions reduced the previously described Kv2.1-mediated enhancement of LTCC activity and CICR in heterologous cells and in cells expressing endogenous channels. These interventions also abolished depolarization-induced and Ca2+-dependent activation of the transcription factors CREB and c-Fos in hippocampal pyramidal neurons [96], showing that the Kv2.1–LTCC ion channel partnership plays an important role in supporting the Ca2+ signaling that underlies excitation–transcription coupling. The effects of these CCAD-based treatments mimicked and occluded the impact of pharmacologically inhibiting LTCC activity, showing that the Kv2.1-mediated regulation of LTCC activity and its localization at ER–PM junctions was crucial to this form of plasticity. As the association of Kv2.1 with LTCCs is specific to neuronal somata, these results provide further experimental support for the privileged role of somatic LTCCs as supplying the bulk of Ca2+ entry leading to depolarization-induced transcript factor activation, as demonstrated in a previous study employing microfluidic chambers to separately activate LTCCs in different neuronal compartments [99]. Together, these studies reveal a unique partnership between a VGCC and a Kv channel mediated through their physical association that impacts the function of the Ca2+ channel and its role in downstream Ca2+ signaling.

4. Restricted Subcellular Compartmentalization of Ion Channel Partnerships

Many ion channels have a highly compartmentalized localization, being found at high densities at specific subcellular sites in neurons [10,67]. As such, the trafficking and retention mechanisms that concentrate ion channels at specific sites in neurons could shape the interactions that underlie selective pairings of specific ion channel partners that underlie the physiologically important partnerships described here. The elevated concentrations of certain candidate ion channel partners at specific sites could be sufficient to provide an environment for selective physical and functional coupling between these ion channels and not others. While the precise subcellular localization of many of the ion channel partnerships detailed above as existing in brain neurons has not been determined, it is clear from biochemical data from brain samples that there exists a direct physical association leading to efficient functional coupling, for example between BK channels and either Cav1-containing VGCCs [55] or Cav2-containing VGCCs [56]. However, there are examples where the partner ion channels have been shown to be selectively co-concentrated at specific sites in brain neurons. In certain cases, we can attribute their colocalization at specific sites to the targeting information present on one of the partners, with the second partner recruited to these sites through this partnership. One example of this scenario is the coclustering of Kv2.1 K+ channels and LTCCs at ER–PM junctions on neuronal somata (Figure 1). The primary driver of the coclustering of LTCCs and Kv2.1 at these sites is the PRC motif present on Kv2.1 [94], which through its binding to ER VAP proteins [91,92] not only localizes Kv2.1 to high-density clusters at these sites but also provides the underlying mechanism for forming and/or maintaining the ER–PM junctions themselves [90,93]. Through their interaction with Kv2.1, which occurs through a mechanism requiring the CCAD motif on Kv2.1 [96], LTCCs concentrate at these ER–PM junctions by associating directly or indirectly with Kv2.1 present at these sites. As detailed above, a prominent consequence of this association is to bring the PM LTCCs into close proximity with ER RyRs facilitating CICR [40]. The Kv2.1–LTCC partnership at these sites may also recruit the Ca2+ channels to the neighborhood of other Ca2+ signaling proteins that are found at these sites [40] and contribute to the privileged role of LTCCs in the Ca2+ entry that underlies excitation–transcription coupling. It is intriguing that many of the other ion channel partnerships detailed above also occur at ER–PM junctions, for example the pairing of ER Ca2+ release channels and PM BK channels in Purkinje cell somata [26], and the tripartite complex of LTCCs, RyRs and BK channels in cartwheel cells [33].
Another example of a highly localized neuronal ion channel partnership is that of T-type (Cav3) Ca2+ channels with KCa3.1 Ca2+-activated K+ channels at cerebellar Purkinje cell axon nodes of Ranvier [100]. The functional association of Ca2+ entry mediated by T-type Ca2+ channels to activate KCa3.1 K+ channels contributes to action potential repolarization that sustains the excitability mediated by the high densities of voltage-gated Na+ channels at these sites. While colocalization of the specific ion channel partners mediating this functional association has not been reported, it is likely the partners are present at the exposed node of Ranvier itself and not in other nodal domains (the paranode and juxtaparanode) where they would be beneath the myelin sheath. It is interesting that BK channels have been found localized to the paranodal region of Purkinje cell axons but not at paranodes on myelinated axons of other brain neurons [101]. Exposure of myelinated Purkinje cell axons to BK channel blockers results in an increase in action potential failure at high firing frequencies [101], suggesting that although they are located under the myelin, BK channels can contribute to axonal function after their activation by an as yet unknown Ca2+ source.
A prominent example of ion channel partnerships that are defined by the respective subcellular localization of their partners are those functioning in dendritic spines of hippocampal CA1 neurons (Figure 2C) [71]. Within this already highly restricted subcellular compartment, further compartmentalization of NMDA receptors and KCa2.2 channels [74] in the postsynaptic density (PSD), and Cav2.3 VGCCs [75] and Kv4.2 channels [76,102] at sites in the spine membrane outside of the PSD define their respective functional coupling. While the NMDA receptor contains a PDZ binding motif mediating its localization within the PSD [103], the molecular mechanism restricting KCa2.2 channels to this subdomain, or Cav2.3 and Kv4 channels to sites within the spine but outside of the PSD itself, is not known. Unlike NMDA receptor subunits, KCa2.2 does not contain a PDZ binding motif to direct its localization to the PSD. However, that the PSD localization is selectively seen for the long but not short KCa2.2 splicing isoform [104] suggests that motifs within the domain that distinguishes the two are required for localization to the PSD, and a subsequent study supports that KCa2.2 localization to the PSD involves the scaffolding protein MPP2 [105]. Whether any of the other ion channel partnerships detailed above are located at distinct sites in neurons, and how the selective targeting of candidate ion channel partners to specific membrane compartments defines the partnerships that have been identified in neurons remains an interesting topic for future research.

5. Conclusions and Future Perspectives

The canonical ion channel partnerships described here involve a specific Ca2+ channel in a physical and functional association with a specific Ca2+-activated K+ channel. However, there also exist non-canonical ion channel partnerships, such as those between VGCCs and voltage-activated Kv4 K+ channels, which are not intrinsically Ca2+-activated but onto which KChIP auxiliary subunits confer Ca2+-dependent modulation. There can also be non-canonical ion channel partnerships wherein the functional relationships between the partners appear to be solely based on their physical association, as in the example of the LTCC–Kv2.1 partnership. Kv2.1-mediated clustering of LTCCs at ER–PM junctions enhances their clustering and activity and increases their spatial and functional association with ER RyRs. There are not any apparent effects on Kv2.1, whose activity is not intrinsically Ca2+-dependent. However, long-term elevation of intracellular [Ca2+] can result in calcineurin-dependent dephosphorylation of Kv2.1, impacting its voltage activation properties and eliminating its clustering at ER–PM junctions [106,107]. This could provide the LTCC–Kv2.1 partnership with a homeostatic mechanism to dynamically reverse the Kv2.1-mediated enhancement of LTCC activity and its association with RyRs under conditions when intracellular [Ca2+] is elevated. The ion channel partnerships described above typically involve the Ca2+ channel partner in an instructive role directing the function of a K+ channel, with the roles being reversed in the Kv2.1:LTCC partnership. A direct reciprocal impact of the K+ channel on its partner Ca2+ channel for the partnerships detailed above have not been described, outside of the effects mediated by changes in membrane potential. Future research may identify bidirectional functional impacts in these partnerships or novel partnerships in which the ion channel partners have reciprocal effects on one another.
It should be noted that while most of the known examples of ion channel partnerships involve Ca2+ channels as one of the partners, there likely exist other forms of neuronal ion channel partnerships. For example, Na+-activated K+ channel activity is coupled to Na+ entry through specific Na+ channel subtypes but not others in brain neurons [108], in somatosensory neurons in dorsal root ganglia [109], and in uterine smooth muscle [110]. This implies close spatial proximity and perhaps physical association between the Na+-activated K+ channel and Na+ channel partners.
The specific molecular mechanisms that generate and maintain the different partners participating in the physical and functional ion channel partnerships reviewed here have, in most cases, not been defined. In many of the examples above, the partnerships defined in native cells have been reconstituted in heterologous expression systems. This allows for structure–function studies defining the domains and then specific amino acids on the ion channel partners that are necessary and sufficient for generating and maintaining the partnership at the physical and/or functional level. This may lead to insights into the potential impact that pathogenic variants in one or the other ion channel partner can have on the partnership, and that could inform genotype-phenotype analyses. It can also lead to the development of specific tools to experimentally intervene in the physical associations underlying the functional consequences of the partnership in native cells. For the LTCC–Kv2.1 partnership, chimeric and mutant isoforms of Kv2.1 were used to define the CCAD motif mediating Kv2.1 association with the Cav1.2 LTCC, leading to the development of point mutants and a cell-penetrating interfering peptide that were then used to selectively disrupt the Kv2.1:Cav1.2 association in native cells [96]. In addition to defining the molecular features that contribute to stable interactions between ion channels, another important area of future investigation will be to identify mechanisms regulating spatiotemporal relationships between ion channel partners. The examples of circadian modulation of Ca2+ source coupling for BK channels in neurons of the suprachiasmatic nucleus [60,111] and neuronal activity-dependent spatial and functional coupling of Kv2.1 and Cav1.2 channels with RyRs in hippocampal CA1 pyramidal neurons [40] demonstrates that ion channel partnerships can be dynamic and reversible. Future studies may reveal molecular mechanisms and spatiotemporal properties of other ion channel partnerships that could lead to further insights into the physiological roles of these partnerships. Understanding these interactions may also provide proof of concept for strategies aimed at disrupting or modulating ion channel partnerships for therapeutic benefit.
Future studies may also reveal other non-canonical pairings, such as the LTCC–Kv2.1 partnership, whose existence could not be predicted on a physiological basis alone. The initial approach that suggested a possible association between Kv2.1 and LTCCs in neurons was an unbiased mass spectrometry-based discovery approach that led to identification of proteins specifically copurifying with Kv2.1 from crosslinked brain samples [40]. Such an approach, which identifies proteins in sufficiently close proximity to be crosslinked to one another, could be used to copurify other ion channels that could be experimentally interrogated for spatial and functional association in heterologous cells and in native cells and tissues. There are numerous proximity labeling approaches [112] that could also be employed, although most of these would require exogenous expression of an ion channel fused to an enzyme capable of mediating the proximity labeling reaction. It is also possible to label cell surface proteins in close proximity to native ion channels using antibody-mediated proximity labeling [113], although to our knowledge, this has not yet been applied to ion channels. Another related approach which has recently yielded important information about the Cav2.3–Kv4.2 partnership is the combination of tandem affinity purification with mass spectrometry [79], which revealed Cav2.3 as a component of Kv4.2-containing protein complexes. Such approaches could lead to identification of other ion channel partnerships, and how they shape physiology above and beyond the respective roles of the individual ion channels and their impact on one another through their effects on membrane potential.

Author Contributions

Conceptualization, N.C.V. and J.S.T.; writing—original draft preparation, N.C.V. and J.S.T.; writing—review and editing, N.C.V. and J.S.T.; funding acquisition, N.C.V. and J.S.T. All authors have read and agreed to the published version of the manuscript.

Funding

The research in our laboratory reviewed here was funded by National Institutes of Health Grants R01NS114210 and R01HL144071 (J.S.T.) and F32NS108519 (N.C.V.).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Dai, S.; Hall, D.D.; Hell, J.W. Supramolecular assemblies and localized regulation of voltage-gated ion channels. Physiol. Rev. 2009, 89, 411–452. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Berridge, M.J. Neuronal calcium signaling. Neuron 1998, 21, 13–26. [Google Scholar] [CrossRef] [Green Version]
  3. Clapham, D.E. Calcium signaling. Cell 2007, 131, 1047–1058. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Augustine, G.J.; Santamaria, F.; Tanaka, K. Local calcium signaling in neurons. Neuron 2003, 40, 331–346. [Google Scholar] [CrossRef] [Green Version]
  5. Neher, E. Vesicle pools and Ca2+ microdomains: New tools for understanding their roles in neurotransmitter release. Neuron 1998, 20, 389–399. [Google Scholar] [CrossRef] [Green Version]
  6. Rizzuto, R.; Pozzan, T. Microdomains of intracellular Ca2+: Molecular determinants and functional consequences. Physiol. Rev. 2006, 86, 369–408. [Google Scholar] [CrossRef]
  7. Grienberger, C.; Konnerth, A. Imaging calcium in neurons. Neuron 2012, 73, 862–885. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Catterall, W.A. Voltage-gated calcium channels. Cold Spring Harb. Perspect. Biol. 2011, 3, a003947. [Google Scholar] [CrossRef]
  9. Simms, B.A.; Zamponi, G.W. Neuronal voltage-gated calcium channels: Structure, function, and dysfunction. Neuron 2014, 82, 24–45. [Google Scholar] [CrossRef] [Green Version]
  10. Vacher, H.; Mohapatra, D.P.; Trimmer, J.S. Localization and targeting of voltage-dependent ion channels in mammalian central neurons. Physiol. Rev. 2008, 88, 1407–1447. [Google Scholar] [CrossRef] [Green Version]
  11. Hell, J.W.; Westenbroek, R.E.; Warner, C.; Ahlijanian, M.K.; Prystay, W.; Gilbert, M.M.; Snutch, T.P.; Catterall, W.A. Identification and differential subcellular localization of the neuronal class C and class D L-type calcium channel alpha 1 subunits. J. Cell Biol. 1993, 123, 949–962. [Google Scholar] [CrossRef]
  12. Obermair, G.J.; Szabo, Z.; Bourinet, E.; Flucher, B.E. Differential targeting of the L-type Ca2+ channel alpha 1C (CaV1.2) to synaptic and extrasynaptic compartments in hippocampal neurons. Eur. J. Neurosci. 2004, 19, 2109–2122. [Google Scholar] [CrossRef] [PubMed]
  13. Leitch, B.; Szostek, A.; Lin, R.; Shevtsova, O. Subcellular distribution of L-type calcium channel subtypes in rat hippocampal neurons. Neuroscience 2009, 164, 641–657. [Google Scholar] [CrossRef]
  14. Stanika, R.I.; Flucher, B.E.; Obermair, G.J. Regulation of postsynaptic stability by the L-type calcium channel Cav1.3 and its interaction with PDZ proteins. Curr. Mol. Pharmacol 2015, 8, 95–101. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Davare, M.A.; Avdonin, V.; Hall, D.D.; Peden, E.M.; Burette, A.; Weinberg, R.J.; Horne, M.C.; Hoshi, T.; Hell, J.W. A beta2 adrenergic receptor signaling complex assembled with the Ca2+ channel Cav1.2. Science 2001, 293, 98–101. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Maximov, A.; Bezprozvanny, I. Synaptic targeting of N-type calcium channels in hippocampal neurons. J. Neurosci. 2002, 22, 6939–6952. [Google Scholar] [CrossRef] [Green Version]
  17. Mochida, S.; Westenbroek, R.E.; Yokoyama, C.T.; Itoh, K.; Catterall, W.A. Subtype-selective reconstitution of synaptic transmission in sympathetic ganglion neurons by expression of exogenous calcium channels. Proc. Natl. Acad. Sci. USA 2003, 100, 2813–2818. [Google Scholar] [CrossRef] [Green Version]
  18. Lory, P.; Nicole, S.; Monteil, A. Neuronal Cav3 channelopathies: Recent progress and perspectives. Pflugers Arch. 2020, 472, 831–844. [Google Scholar] [CrossRef]
  19. Striessnig, J. Voltage-gated Ca2+-channel alpha1-subunit de novo missense mutations: Gain or loss of function—implications for potential therapies. Front. Synaptic Neurosci. 2021, 13, 634760. [Google Scholar] [CrossRef]
  20. Eggermann, E.; Bucurenciu, I.; Goswami, S.P.; Jonas, P. Nanodomain coupling between Ca2+ channels and sensors of exocytosis at fast mammalian synapses. Nat. Rev. Neurosci. 2011, 13, 7–21. [Google Scholar] [CrossRef]
  21. Dixon, R.E.; Navedo, M.F.; Binder, M.D.; Santana, L.F. Mechanisms and physiological implications of cooperative gating of ion channels clusters. Physiol. Rev. 2021; online ahead of print. [Google Scholar] [CrossRef]
  22. Dixon, R.E.; Yuan, C.; Cheng, E.P.; Navedo, M.F.; Santana, L.F. Ca2+ signaling amplification by oligomerization of L-type Cav1.2 channels. Proc. Natl. Acad. Sci. USA 2012, 109, 1749–1754. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Dixon, R.E.; Moreno, C.M.; Yuan, C.; Opitz-Araya, X.; Binder, M.D.; Navedo, M.F.; Santana, L.F. Graded Ca2+/calmodulin-dependent coupling of voltage-gated Cav1.2 channels. eLife 2015, 4, e05608. [Google Scholar] [CrossRef] [PubMed]
  24. Rosenbluth, J. Subsurface cisterns and their relationship to the neuronal plasma membrane. J. Cell Biol. 1962, 13, 405–421. [Google Scholar] [CrossRef] [PubMed]
  25. Wu, Y.; Whiteus, C.; Xu, C.S.; Hayworth, K.J.; Weinberg, R.J.; Hess, H.F.; De Camilli, P. Contacts between the endoplasmic reticulum and other membranes in neurons. Proc. Natl. Acad. Sci. USA 2017, 114, E4859–E4867. [Google Scholar] [CrossRef] [Green Version]
  26. Kaufmann, W.A.; Ferraguti, F.; Fukazawa, Y.; Kasugai, Y.; Shigemoto, R.; Laake, P.; Sexton, J.A.; Ruth, P.; Wietzorrek, G.; Knaus, H.G.; et al. Large-conductance calcium-activated potassium channels in purkinje cell plasma membranes are clustered at sites of hypolemmal microdomains. J. Comp. Neurol. 2009, 515, 215–230. [Google Scholar] [CrossRef]
  27. Mandikian, D.; Bocksteins, E.; Parajuli, L.K.; Bishop, H.I.; Cerda, O.; Shigemoto, R.; Trimmer, J.S. Cell type-specific spatial and functional coupling between mammalian brain Kv2.1 K+ channels and ryanodine receptors. J. Comp. Neurol. 2014, 522, 3555–3574. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Arias-Cavieres, A.; Barrientos, G.C.; Sanchez, G.; Elgueta, C.; Munoz, P.; Hidalgo, C. Ryanodine receptor-mediated calcium release has a key role in hippocampal LTD induction. Front. Cell Neurosci. 2018, 12, 403. [Google Scholar] [CrossRef] [Green Version]
  29. Lobos, P.; Cordova, A.; Vega-Vasquez, I.; Ramirez, O.A.; Adasme, T.; Toledo, J.; Cerda, M.; Hartel, S.; Paula-Lima, A.; Hidalgo, C. RyR-mediated Ca2+ release elicited by neuronal activity induces nuclear Ca2+ signals, CREB phosphorylation, and Npas4/RyR2 expression. Proc. Natl. Acad. Sci. USA 2021, 118, e2102265118. [Google Scholar] [CrossRef]
  30. Bertan, F.; Wischhof, L.; Sosulina, L.; Mittag, M.; Dalugge, D.; Fornarelli, A.; Gardoni, F.; Marcello, E.; Di Luca, M.; Fuhrmann, M.; et al. Loss of Ryanodine Receptor 2 impairs neuronal activity-dependent remodeling of dendritic spines and triggers compensatory neuronal hyperexcitability. Cell Death Differ. 2020, 27, 3354–3373. [Google Scholar] [CrossRef]
  31. Sah, P.; McLachlan, E.M. Ca2+-activated K+ currents underlying the afterhyperpolarization in guinea pig vagal neurons: A role for Ca2+-activated Ca2+ release. Neuron 1991, 7, 257–264. [Google Scholar] [CrossRef]
  32. Wang, B.; Bugay, V.; Ling, L.; Chuang, H.H.; Jaffe, D.B.; Brenner, R. Knockout of the BK beta4-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] [PubMed] [Green Version]
  33. 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] [PubMed] [Green Version]
  34. Cui, G.; Okamoto, T.; Morikawa, H. Spontaneous opening of T-type Ca2+ channels contributes to the irregular firing of dopamine neurons in neonatal rats. J. Neurosci. 2004, 24, 11079–11087. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Lipkin, A.M.; Cunniff, M.M.; Spratt, P.W.E.; Lemke, S.M.; Bender, K.J. Functional microstructure of Cav-mediated calcium signaling in the axon initial segment. J. Neurosci. 2021, 41, 3764–3776. [Google Scholar] [CrossRef] [PubMed]
  36. Wang, S.Q.; Song, L.S.; Lakatta, E.G.; Cheng, H. Ca2+ signalling between single L-type Ca2+ channels and ryanodine receptors in heart cells. Nature 2001, 410, 592–596. [Google Scholar] [CrossRef] [PubMed]
  37. Cheng, H.; Lederer, W.J. Calcium sparks. Physiol. Rev. 2008, 88, 1491–1545. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Manita, S.; Ross, W.N. Synaptic activation and membrane potential changes modulate the frequency of spontaneous elementary Ca2+ release events in the dendrites of pyramidal neurons. J. Neurosci. 2009, 29, 7833–7845. [Google Scholar] [CrossRef] [Green Version]
  39. Berrout, J.; Isokawa, M. Homeostatic and stimulus-induced coupling of the L-type Ca2+ channel to the ryanodine receptor in the hippocampal neuron in slices. Cell Calcium 2009, 46, 30–38. [Google Scholar] [CrossRef] [Green Version]
  40. Vierra, N.C.; Kirmiz, M.; van der List, D.; Santana, L.F.; Trimmer, J.S. Kv2.1 mediates spatial and functional coupling of L-type calcium channels and ryanodine receptors in mammalian neurons. eLife 2019, 8, e49953. [Google Scholar] [CrossRef]
  41. Stern, M.D.; Song, L.S.; Cheng, H.; Sham, J.S.; Yang, H.T.; Boheler, K.R.; Rios, E. Local control models of cardiac excitation-contraction coupling. A possible role for allosteric interactions between ryanodine receptors. J. Gen. Physiol. 1999, 113, 469–489. [Google Scholar] [CrossRef] [Green Version]
  42. Jacobs, J.M.; Meyer, T. Control of action potential-induced Ca2+ signaling in the soma of hippocampal neurons by Ca2+ release from intracellular stores. J. Neurosci. 1997, 17, 4129–4135. [Google Scholar] [CrossRef] [Green Version]
  43. Dittmer, P.J.; Wild, A.R.; Dell’Acqua, M.L.; Sather, W.A. STIM1 Ca2+ sensor control of L-type Ca2+-channel-dependent dendritic spine structural plasticity and nuclear signaling. Cell Rep. 2017, 19, 321–334. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Dittmer, P.J.; Dell’Acqua, M.L.; Sather, W.A. Synaptic crosstalk conferred by a zone of differentially regulated Ca2+ signaling in the dendritic shaft adjoining a potentiated spine. Proc. Natl. Acad Sci. USA 2019, 116, 13611–13620. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Sather, W.A.; Dittmer, P.J. Regulation of voltage-gated calcium channels by the ER calcium sensor STIM1. Curr. Opin. Neurobiol. 2019, 57, 186–191. [Google Scholar] [CrossRef] [PubMed]
  46. Contet, C.; Goulding, S.P.; Kuljis, D.A.; Barth, A.L. BK channels in the central nervous system. Int. Rev. Neurobiol. 2016, 128, 281–342. [Google Scholar]
  47. Adelman, J.P.; Maylie, J.; Sah, P. Small-conductance Ca2+-activated K+ channels: Form and function. Annu. Rev. Physiol. 2012, 74, 245–269. [Google Scholar] [CrossRef]
  48. Kakizawa, S.; Kishimoto, Y.; Hashimoto, K.; Miyazaki, T.; Furutani, K.; Shimizu, H.; Fukaya, M.; Nishi, M.; Sakagami, H.; Ikeda, A.; et al. Junctophilin-mediated channel crosstalk essential for cerebellar synaptic plasticity. EMBO J. 2007, 26, 1924–1933. [Google Scholar] [CrossRef] [Green Version]
  49. Fakler, B.; Adelman, J.P. Control of K(Ca) channels by calcium nano/microdomains. Neuron 2008, 59, 873–881. [Google Scholar] [CrossRef] [Green Version]
  50. Kaufmann, W.A.; Kasugai, Y.; Ferraguti, F.; Storm, J.F. 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] [Green Version]
  51. Indriati, D.W.; Kamasawa, N.; Matsui, K.; Meredith, A.L.; Watanabe, M.; Shigemoto, R. Quantitative localization of Cav2.1 (P/Q-type) voltage-dependent calcium channels in Purkinje cells: Somatodendritic gradient and distinct somatic coclustering with calcium-activated potassium channels. J. Neurosci. 2013, 33, 3668–3678. [Google Scholar] [CrossRef] [Green Version]
  52. Lancaster, B.; Nicoll, R.A. Properties of two calcium-activated hyperpolarizations in rat hippocampal neurones. J. Physiol. 1987, 389, 187–203. [Google Scholar] [CrossRef] [PubMed]
  53. 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] [PubMed]
  54. Berkefeld, H.; Fakler, B.; Schulte, U. Ca2+-activated K+ channels: From protein complexes to function. Physiol. Rev. 2010, 90, 1437–1459. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. 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 Ca2+-activated K+ signaling. Science 2006, 314, 615–620. [Google Scholar] [CrossRef] [Green Version]
  56. Muller, C.S.; Haupt, A.; Bildl, W.; Schindler, J.; Knaus, H.G.; Meissner, M.; Rammner, B.; Striessnig, J.; Flockerzi, V.; Fakler, B.; et al. Quantitative proteomics of the Cav2 channel nano-environments in the mammalian brain. Proc. Natl. Acad. Sci. USA 2010, 107, 14950–14957. [Google Scholar] [CrossRef] [Green Version]
  57. Yazejian, B.; DiGregorio, D.A.; Vergara, J.L.; Poage, R.E.; Meriney, S.D.; Grinnell, A.D. Direct measurements of presynaptic calcium and calcium-activated potassium currents regulating neurotransmitter release at cultured Xenopus nerve-muscle synapses. J. Neurosci. 1997, 17, 2990–3001. [Google Scholar] [CrossRef] [Green Version]
  58. 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]
  59. 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]
  60. 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. 2018, 150, 259–275. [Google Scholar] [CrossRef] [Green Version]
  61. Schumacher, M.A.; Rivard, A.F.; Bachinger, H.P.; Adelman, J.P. Structure of the gating domain of a Ca2+-activated K+ channel complexed with Ca2+/calmodulin. Nature 2001, 410, 1120–1124. [Google Scholar] [CrossRef]
  62. King, B.; Rizwan, A.P.; Asmara, H.; Heath, N.C.; Engbers, J.D.; Dykstra, S.; Bartoletti, T.M.; Hameed, S.; Zamponi, G.W.; Turner, R.W. IKCa channels are a critical determinant of the slow AHP in CA1 pyramidal neurons. Cell Rep. 2015, 11, 175–182. [Google Scholar] [CrossRef] [Green Version]
  63. Wang, K.; Mateos-Aparicio, P.; Honigsperger, C.; Raghuram, V.; Wu, W.W.; Ridder, M.C.; Sah, P.; Maylie, J.; Storm, J.F.; Adelman, J.P. IK1 channels do not contribute to the slow afterhyperpolarization in pyramidal neurons. eLife 2016, 5, e11206. [Google Scholar] [CrossRef] [Green Version]
  64. Sahu, G.; Asmara, H.; Zhang, F.X.; Zamponi, G.W.; Turner, R.W. Activity-dependent facilitation of Cav1.3 calcium channels promotes KCa3.1 activation in hippocampal neurons. J. Neurosci 2017, 37, 11255–11270. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Sahu, G.; Wazen, R.M.; Colarusso, P.; Chen, S.R.W.; Zamponi, G.W.; Turner, R.W. Junctophilin proteins tether a Cav1-RyR2-KCa3.1 tripartite complex to regulate neuronal excitability. Cell Rep. 2019, 28, 2427–2442.e6. [Google Scholar] [CrossRef] [Green Version]
  66. Anderson, D.; Mehaffey, W.H.; Iftinca, M.; Rehak, R.; Engbers, J.D.; Hameed, S.; Zamponi, G.W.; Turner, R.W. Regulation of neuronal activity by Cav3-Kv4 channel signaling complexes. Nat. Neurosci. 2010, 13, 333–337. [Google Scholar] [CrossRef]
  67. Trimmer, J.S. Subcellular localization of K+ channels in mammalian brain neurons: Remarkable precision in the midst of extraordinary complexity. Neuron 2015, 85, 238–256. [Google Scholar] [CrossRef] [Green Version]
  68. An, W.F.; Bowlby, M.R.; Betty, M.; Cao, J.; Ling, H.P.; Mendoza, G.; Hinson, J.W.; Mattsson, K.I.; Strassle, B.W.; Trimmer, J.S.; et al. Modulation of A-type potassium channels by a family of calcium sensors. Nature 2000, 403, 553–556. [Google Scholar] [CrossRef] [PubMed]
  69. Rhodes, K.J.; Carroll, K.I.; Sung, M.A.; Doliveira, L.C.; Monaghan, M.M.; Burke, S.L.; Strassle, B.W.; Buchwalder, L.; Menegola, M.; Cao, J.; et al. KChIPs and Kv4 alpha subunits as integral components of A-type potassium channels in mammalian brain. J. Neurosci. 2004, 24, 7903–7915. [Google Scholar] [CrossRef] [PubMed]
  70. Anderson, D.; Engbers, J.D.; Heath, N.C.; Bartoletti, T.M.; Mehaffey, W.H.; Zamponi, G.W.; Turner, R.W. The Cav3-Kv4 complex acts as a calcium sensor to maintain inhibitory charge transfer during extracellular calcium fluctuations. J. Neurosci. 2013, 33, 7811–7824. [Google Scholar] [CrossRef] [Green Version]
  71. Wang, K.; Lin, M.T.; Adelman, J.P.; Maylie, J. Distinct Ca2+ sources in dendritic spines of hippocampal CA1 neurons couple to SK and Kv4 channels. Neuron 2014, 81, 379–387. [Google Scholar] [CrossRef] [Green Version]
  72. Kimm, T.; Bean, B.P. Inhibition of A-type potassium current by the peptide toxin SNX-482. J. Neurosci 2014, 34, 9182–9189. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Ngo-Anh, T.J.; Bloodgood, B.L.; Lin, M.; Sabatini, B.L.; Maylie, J.; Adelman, J.P. SK channels and NMDA receptors form a Ca2+-mediated feedback loop in dendritic spines. Nat. Neurosci. 2005, 8, 642–649. [Google Scholar] [CrossRef] [PubMed]
  74. Lin, M.T.; Lujan, R.; Watanabe, M.; Adelman, J.P.; Maylie, J. SK2 channel plasticity contributes to LTP at Schaffer collateral-CA1 synapses. Nat. Neurosci. 2008, 11, 170–177. [Google Scholar] [CrossRef] [PubMed]
  75. Parajuli, L.K.; Nakajima, C.; Kulik, A.; Matsui, K.; Schneider, T.; Shigemoto, R.; Fukazawa, Y. Quantitative regional and ultrastructural localization of the Ca(v)2.3 subunit of R-type calcium channel in mouse brain. J. Neurosci. 2012, 32, 13555–13567. [Google Scholar] [CrossRef] [Green Version]
  76. Kim, J.; Jung, S.C.; Clemens, A.M.; Petralia, R.S.; Hoffman, D.A. Regulation of dendritic excitability by activity-dependent trafficking of the A-type K+ channel subunit Kv4.2 in hippocampal neurons. Neuron 2007, 54, 933–947. [Google Scholar] [CrossRef] [Green Version]
  77. Wang, K.; Kelley, M.H.; Wu, W.W.; Adelman, J.P.; Maylie, J. Apamin boosting of synaptic potentials in Cav2.3 R-type Ca2+ channel null mice. PLoS ONE 2015, 10, e0139332. [Google Scholar] [CrossRef]
  78. Giessel, A.J.; Sabatini, B.L. Boosting of synaptic potentials and spine Ca transients by the peptide toxin SNX-482 requires alpha-1E-encoded voltage-gated Ca channels. PLoS ONE 2011, 6, e20939. [Google Scholar] [CrossRef]
  79. Murphy, J.G.; Gutzmann, J.J.; Lin, L.; Hu, J.; Petralia, R.S.; Wang, Y.X.; Hoffman, D.A. R-type voltage-gated Ca2+ channels mediate A-type K+ current regulation of synaptic input in hippocampal dendrites. Cell Rep. 2022, 38, 110264. [Google Scholar] [CrossRef]
  80. Hoffman, D.A.; Magee, J.C.; Colbert, C.M.; Johnston, D. K+ channel regulation of signal propagation in dendrites of hippocampal pyramidal neurons. Nature 1997, 387, 869–875. [Google Scholar] [CrossRef]
  81. Heath, N.C.; Rizwan, A.P.; Engbers, J.D.; Anderson, D.; Zamponi, G.W.; Turner, R.W. The expression pattern of a Cav3-Kv4 complex differentially regulates spike output in cerebellar granule cells. J. Neurosci. 2014, 34, 8800–8812. [Google Scholar] [CrossRef] [Green Version]
  82. Serodio, P.; Rudy, B. Differential expression of Kv4 K+ channel subunits mediating subthreshold transient K+ (A-type) currents in rat brain. J. Neurophysiol. 1998, 79, 1081–1091. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Strassle, B.W.; Menegola, M.; Rhodes, K.J.; Trimmer, J.S. Light and electron microscopic analysis of KChIP and Kv4 localization in rat cerebellar granule cells. J. Comp. Neurol. 2005, 484, 144–155. [Google Scholar] [CrossRef] [PubMed]
  84. Amarillo, Y.; De Santiago-Castillo, J.A.; Dougherty, K.; Maffie, J.; Kwon, E.; Covarrubias, M.; Rudy, B. Ternary Kv4.2 channels recapitulate voltage-dependent inactivation kinetics of A-type K+ channels in cerebellar granule neurons. J. Physiol. 2008, 586, 2093–2106. [Google Scholar] [CrossRef]
  85. Dayal, A.; Schrotter, K.; Pan, Y.; Fohr, K.; Melzer, W.; Grabner, M. The Ca2+ influx through the mammalian skeletal muscle dihydropyridine receptor is irrelevant for muscle performance. Nat. Commun. 2017, 8, 475. [Google Scholar] [CrossRef] [Green Version]
  86. Du, J.; Tao-Cheng, J.H.; Zerfas, P.; McBain, C.J. The K+ channel, Kv2.1, is apposed to astrocytic processes and is associated with inhibitory postsynaptic membranes in hippocampal and cortical principal neurons and inhibitory interneurons. Neuroscience 1998, 84, 37–48. [Google Scholar] [CrossRef]
  87. Antonucci, D.E.; Lim, S.T.; Vassanelli, S.; Trimmer, J.S. Dynamic localization and clustering of dendritic Kv2.1 voltage-dependent potassium channels in developing hippocampal neurons. Neuroscience 2001, 108, 69–81. [Google Scholar] [CrossRef]
  88. Misonou, H.; Mohapatra, D.P.; Trimmer, J.S. Kv2.1: A voltage-gated K+ channel critical to dynamic control of neuronal excitability. Neurotoxicology 2005, 26, 743–752. [Google Scholar] [CrossRef]
  89. King, A.N.; Manning, C.F.; Trimmer, J.S. A unique ion channel clustering domain on the axon initial segment of mammalian neurons. J. Comp. Neurol. 2014, 522, 2594–2608. [Google Scholar] [CrossRef] [Green Version]
  90. Fox, P.D.; Haberkorn, C.J.; Akin, E.J.; Seel, P.J.; Krapf, D.; Tamkun, M.M. Induction of stable ER-plasma-membrane junctions by Kv2.1 potassium channels. J. Cell Sci. 2015, 128, 2096–2105. [Google Scholar] [CrossRef] [Green Version]
  91. Johnson, B.; Leek, A.N.; Sole, L.; Maverick, E.E.; Levine, T.P.; Tamkun, M.M. Kv2 potassium channels form endoplasmic reticulum/plasma membrane junctions via interaction with VAPA and VAPB. Proc. Natl. Acad. Sci. USA 2018, 115, E7331–E7340. [Google Scholar] [CrossRef] [Green Version]
  92. Kirmiz, M.; Vierra, N.C.; Palacio, S.; Trimmer, J.S. Identification of VAPA and VAPB as Kv2 channel-interacting proteins defining endoplasmic reticulum-plasma membrane junctions in mammalian brain neurons. J. Neurosci 2018, 38, 7562–7584. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Kirmiz, M.; Palacio, S.; Thapa, P.; King, A.N.; Sack, J.T.; Trimmer, J.S. Remodeling neuronal ER-PM junctions is a conserved nonconducting function of Kv2 plasma membrane ion channels. Mol. Biol Cell 2018, 29, 2410–2432. [Google Scholar] [CrossRef]
  94. Lim, S.T.; Antonucci, D.E.; Scannevin, R.H.; Trimmer, J.S. A novel targeting signal for proximal clustering of the Kv2.1 K+ channel in hippocampal neurons. Neuron 2000, 25, 385–397. [Google Scholar] [CrossRef] [Green Version]
  95. O’Dwyer, S.C.; Palacio, S.; Matsumoto, C.; Guarina, L.; Klug, N.R.; Tajada, S.; Rosati, B.; McKinnon, D.; Trimmer, J.S.; Santana, L.F. Kv2.1 channels play opposing roles in regulating membrane potential, Ca2+ channel function, and myogenic tone in arterial smooth muscle. Proc. Natl. Acad. Sci. USA 2020, 117, 3858–3866. [Google Scholar] [CrossRef] [Green Version]
  96. Vierra, N.C.; O’Dwyer, S.C.; Matsumoto, C.; Santana, L.F.; Trimmer, J.S. Regulation of neuronal excitation-transcription coupling by Kv2.1-induced clustering of somatic L-type Ca2+ channels at ER-PM junctions. Proc. Natl. Acad. Sci. USA 2021, 118, e2110094118. [Google Scholar] [CrossRef] [PubMed]
  97. Fredriksson, S.; Gullberg, M.; Jarvius, J.; Olsson, C.; Pietras, K.; Gustafsdottir, S.M.; Ostman, A.; Landegren, U. Protein detection using proximity-dependent DNA ligation assays. Nat. Biotechnol. 2002, 20, 473–477. [Google Scholar] [CrossRef]
  98. Soderberg, O.; Gullberg, M.; Jarvius, M.; Ridderstrale, K.; Leuchowius, K.J.; Jarvius, J.; Wester, K.; Hydbring, P.; Bahram, F.; Larsson, L.G.; et al. Direct observation of individual endogenous protein complexes in situ by proximity ligation. Nat. Methods 2006, 3, 995–1000. [Google Scholar] [CrossRef]
  99. Wild, A.R.; Sinnen, B.L.; Dittmer, P.J.; Kennedy, M.J.; Sather, W.A.; Dell’Acqua, M.L. Synapse-to-nucleus communication through NFAT Is mediated by L-type Ca2+ channel Ca2+ spike propagation to the soma. Cell Rep. 2019, 26, 3537–3550.e4. [Google Scholar] [CrossRef] [Green Version]
  100. Grundemann, J.; Clark, B.A. Calcium-activated potassium channels at nodes of Ranvier secure axonal spike propagation. Cell Rep. 2015, 12, 1715–1722. [Google Scholar] [CrossRef] [Green Version]
  101. 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] [Green Version]
  102. Kerti, K.; Lorincz, A.; Nusser, Z. Unique somato-dendritic distribution pattern of Kv4.2 channels on hippocampal CA1 pyramidal cells. Eur. J. Neurosci. 2012, 35, 66–75. [Google Scholar] [CrossRef] [PubMed]
  103. Prybylowski, K.; Chang, K.; Sans, N.; Kan, L.; Vicini, S.; Wenthold, R.J. The synaptic localization of NR2B-containing NMDA receptors is controlled by interactions with PDZ proteins and AP-2. Neuron 2005, 47, 845–857. [Google Scholar] [CrossRef] [Green Version]
  104. Allen, D.; Bond, C.T.; Lujan, R.; Ballesteros-Merino, C.; Lin, M.T.; Wang, K.; Klett, N.; Watanabe, M.; Shigemoto, R.; Stackman, R.W., Jr.; et al. The SK2-long isoform directs synaptic localization and function of SK2-containing channels. Nat. Neurosci. 2011, 14, 744–749. [Google Scholar] [CrossRef] [Green Version]
  105. Kim, G.; Lujan, R.; Schwenk, J.; Kelley, M.H.; Aguado, C.; Watanabe, M.; Fakler, B.; Maylie, J.; Adelman, J.P. Membrane palmitoylated protein 2 is a synaptic scaffold protein.n required for synaptic SK2-containing channel function. eLife 2016, 5, e12637. [Google Scholar] [CrossRef]
  106. Misonou, H.; Mohapatra, D.P.; Park, E.W.; Leung, V.; Zhen, D.; Misonou, K.; Anderson, A.E.; Trimmer, J.S. Regulation of ion channel localization and phosphorylation by neuronal activity. Nat. Neurosci. 2004, 7, 711–718. [Google Scholar] [CrossRef] [PubMed]
  107. Misonou, H.; Menegola, M.; Mohapatra, D.P.; Guy, L.K.; Park, K.S.; Trimmer, J.S. Bidirectional activity-dependent regulation of neuronal ion channel phosphorylation. J. Neurosci. 2006, 26, 13505–13514. [Google Scholar] [CrossRef] [Green Version]
  108. Hage, T.A.; Salkoff, L. Sodium-activated potassium channels are functionally coupled to persistent sodium currents. J. Neurosci. 2012, 32, 2714–2721. [Google Scholar] [CrossRef] [Green Version]
  109. Pryce, K.D.; Powell, R.; Agwa, D.; Evely, K.M.; Sheehan, G.D.; Nip, A.; Tomasello, D.L.; Gururaj, S.; Bhattacharjee, A. Magi-1 scaffolds NaV1.8 and Slack KNa channels in dorsal root ganglion neurons regulating excitability and pain. FASEB J. 2019, 33, 7315–7330. [Google Scholar] [CrossRef] [PubMed]
  110. Ferreira, J.J.; Amazu, C.; Puga-Molina, L.C.; Ma, X.; England, S.K.; Santi, C.M. SLO2.1/NALCN a sodium signaling complex that regulates uterine activity. iScience 2021, 24, 103210. [Google Scholar] [CrossRef]
  111. 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]
  112. Qin, W.; Cho, K.F.; Cavanagh, P.E.; Ting, A.Y. Deciphering molecular interactions by proximity labeling. Nat. Methods 2021, 18, 133–143. [Google Scholar] [CrossRef] [PubMed]
  113. Bar, D.Z.; Atkatsh, K.; Tavarez, U.; Erdos, M.R.; Gruenbaum, Y.; Collins, F.S. Biotinylation by antibody recognition-a method for proximity labeling. Nat. Methods 2018, 15, 127–133. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Spatial and functional coupling of Cav1.2 and RyRs at ER–PM contact sites. The left panel is a schematic of a neuron showing that the soma is the subcellular compartment in which this partnership has been defined. The right panel shows that PM Kv2.1 channels (red) form and stabilize ER–PM junctions via a physical interaction with ER-resident VAPA and VAPB proteins (dark grey). Kv2.1 simultaneously promotes the clustering of PM Cav1.2 channels (cyan) at these same sites, placing Cav1.2 channels near RyRs (dark blue) localized in the ER. This arrangement enables Ca2+ influx through Cav1.2 channels to trigger enhanced RyR-dependent CICR. Intracellular Ca2+ is shown in magenta.
Figure 1. Spatial and functional coupling of Cav1.2 and RyRs at ER–PM contact sites. The left panel is a schematic of a neuron showing that the soma is the subcellular compartment in which this partnership has been defined. The right panel shows that PM Kv2.1 channels (red) form and stabilize ER–PM junctions via a physical interaction with ER-resident VAPA and VAPB proteins (dark grey). Kv2.1 simultaneously promotes the clustering of PM Cav1.2 channels (cyan) at these same sites, placing Cav1.2 channels near RyRs (dark blue) localized in the ER. This arrangement enables Ca2+ influx through Cav1.2 channels to trigger enhanced RyR-dependent CICR. Intracellular Ca2+ is shown in magenta.
Ijms 23 01953 g001
Figure 2. Ca2+-activated BK and SK channels form distinct partnerships with different neuronal Ca2+ channels. (A) Schematic of a neuron showing the subcellular compartments in which the partnerships shown in panels B–D have been defined. (B) PM BK channels can partner with ER RyRs and PM VGCCs. In somatodendritic domains of cartwheel neurons, a partnership of PM Cav2.1 channels (green) and ER RyRs (blue) works together to activate spatially and functionally associated PM BK channels (orange). These BK currents work to control patterns of electrical activity in these cells. (C) PM BK channels (orange) associated with PM Cav2.1 channels (green) in presynaptic terminals (containing schematized synaptic vesicles [SVs]) are activated by Ca2+ influx through these channels. BK current acts to hyperpolarize the membrane potential, limiting Ca2+ influx into the terminal by promoting closure of Cav2 channels. In dendritic spines, the partnership of SK channels (yellow) and NMDA receptors (purple) at or near the PSD (schematized in black), and the partnership of Cav2.3 channels (dark green) and Kv4.2 channels (dark purple) at sites outside of the PSD control EPSP amplitude. Together, these partnerships function to modulate distinct aspects of neurotransmission. (D) Somatodendritic BK channels (orange) can partner with Cav1.3 channels (light green) and also Cav2.3 channels (dark green). Action potential-triggered Ca2+ influx activates BK channels in these partnerships, providing cells with a mechanism to rapidly modulate electrical activity. Intracellular Ca2+ is shown in magenta in all panels BD.
Figure 2. Ca2+-activated BK and SK channels form distinct partnerships with different neuronal Ca2+ channels. (A) Schematic of a neuron showing the subcellular compartments in which the partnerships shown in panels B–D have been defined. (B) PM BK channels can partner with ER RyRs and PM VGCCs. In somatodendritic domains of cartwheel neurons, a partnership of PM Cav2.1 channels (green) and ER RyRs (blue) works together to activate spatially and functionally associated PM BK channels (orange). These BK currents work to control patterns of electrical activity in these cells. (C) PM BK channels (orange) associated with PM Cav2.1 channels (green) in presynaptic terminals (containing schematized synaptic vesicles [SVs]) are activated by Ca2+ influx through these channels. BK current acts to hyperpolarize the membrane potential, limiting Ca2+ influx into the terminal by promoting closure of Cav2 channels. In dendritic spines, the partnership of SK channels (yellow) and NMDA receptors (purple) at or near the PSD (schematized in black), and the partnership of Cav2.3 channels (dark green) and Kv4.2 channels (dark purple) at sites outside of the PSD control EPSP amplitude. Together, these partnerships function to modulate distinct aspects of neurotransmission. (D) Somatodendritic BK channels (orange) can partner with Cav1.3 channels (light green) and also Cav2.3 channels (dark green). Action potential-triggered Ca2+ influx activates BK channels in these partnerships, providing cells with a mechanism to rapidly modulate electrical activity. Intracellular Ca2+ is shown in magenta in all panels BD.
Ijms 23 01953 g002
Figure 3. The partnership of somatic Cav1.3, RyR2, and IK channels at junctophilin-3 and -4 stabilized ER–PM junctions. The left panel is a schematic of a neuron showing that the soma is the subcellular compartment in which this partnership has been defined. The right panel shows a schematic of the partnership defined by Sahu and colleagues, who demonstrated that the spatial and functional coupling of PM Cav1.3 channels (green) with ER RyR2 channels (blue) and PM IK channels (yellow) at somatic ER–PM junctions required the presence of junctophilin-3 (JPH3, purple) and -4 (JPH4, magenta) proteins, which form and stabilize contacts between the ER and PM. Loss of JPH3 and JPH4 reduced the spatial association of Cav1.3, RyR2, and IK channels in hippocampal CA1 pyramidal neurons, impairing the CICR-mediated activation of Ca2+-activated IK channels and altering intrinsic excitability. Intracellular Ca2+ is shown in magenta.
Figure 3. The partnership of somatic Cav1.3, RyR2, and IK channels at junctophilin-3 and -4 stabilized ER–PM junctions. The left panel is a schematic of a neuron showing that the soma is the subcellular compartment in which this partnership has been defined. The right panel shows a schematic of the partnership defined by Sahu and colleagues, who demonstrated that the spatial and functional coupling of PM Cav1.3 channels (green) with ER RyR2 channels (blue) and PM IK channels (yellow) at somatic ER–PM junctions required the presence of junctophilin-3 (JPH3, purple) and -4 (JPH4, magenta) proteins, which form and stabilize contacts between the ER and PM. Loss of JPH3 and JPH4 reduced the spatial association of Cav1.3, RyR2, and IK channels in hippocampal CA1 pyramidal neurons, impairing the CICR-mediated activation of Ca2+-activated IK channels and altering intrinsic excitability. Intracellular Ca2+ is shown in magenta.
Ijms 23 01953 g003
Figure 4. Partnerships of somatodendritic VGCCs with Kv channels. (A) The left panel is a schematic of a neuron showing that the somatodendritic compartment is where this partnership has been defined. The right panel shows Ca2+ influx through PM T-type Cav3 channels (Cav3 channel in blue) can trigger the activation of associated PM Kv4 channels (purple). (B) The left panel is a schematic of a neuron showing that the dendritic compartment is where this partnership has been defined. The right panel shows Ca2+ influx through PM R-type Cav2.3 channels (Cav2.3 channel in green) can trigger the activation of associated PM Kv4 channels (purple). In both of these cases, cytoplasmic KChIP auxiliary subunits (shown as purple rectangles) associated with Kv4 channels function as the Ca2+ sensor that triggers the opening of Kv4 channels. These partnerships confer Ca2+-dependent activation to prominent mediators of IA to dampen electrical excitability. Intracellular Ca2+ is shown in magenta in both right panels.
Figure 4. Partnerships of somatodendritic VGCCs with Kv channels. (A) The left panel is a schematic of a neuron showing that the somatodendritic compartment is where this partnership has been defined. The right panel shows Ca2+ influx through PM T-type Cav3 channels (Cav3 channel in blue) can trigger the activation of associated PM Kv4 channels (purple). (B) The left panel is a schematic of a neuron showing that the dendritic compartment is where this partnership has been defined. The right panel shows Ca2+ influx through PM R-type Cav2.3 channels (Cav2.3 channel in green) can trigger the activation of associated PM Kv4 channels (purple). In both of these cases, cytoplasmic KChIP auxiliary subunits (shown as purple rectangles) associated with Kv4 channels function as the Ca2+ sensor that triggers the opening of Kv4 channels. These partnerships confer Ca2+-dependent activation to prominent mediators of IA to dampen electrical excitability. Intracellular Ca2+ is shown in magenta in both right panels.
Ijms 23 01953 g004
Table 1. VGCCs subtypes.
Table 1. VGCCs subtypes.
Cav α1 SubunitCa2+ Current
Cav1.1L-type
Cav1.2L-type
Cav1.3L-type
Cav1.4L-type
Cav2.1P/Q-type
Cav2.2N-type
Cav2.3R-type
Cav3.1T-type
Cav3.2T-type
Cav3.3T-type
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Vierra, N.C.; Trimmer, J.S. Ion Channel Partnerships: Odd and Not-So-Odd Couples Controlling Neuronal Ion Channel Function. Int. J. Mol. Sci. 2022, 23, 1953. https://doi.org/10.3390/ijms23041953

AMA Style

Vierra NC, Trimmer JS. Ion Channel Partnerships: Odd and Not-So-Odd Couples Controlling Neuronal Ion Channel Function. International Journal of Molecular Sciences. 2022; 23(4):1953. https://doi.org/10.3390/ijms23041953

Chicago/Turabian Style

Vierra, Nicholas C., and James S. Trimmer. 2022. "Ion Channel Partnerships: Odd and Not-So-Odd Couples Controlling Neuronal Ion Channel Function" International Journal of Molecular Sciences 23, no. 4: 1953. https://doi.org/10.3390/ijms23041953

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