Voltage-gated Ca2+ channels (VGCCs) are highly expressed throughout numerous biological systems and play critical roles in synaptic transmission, cardiac excitation, and muscle contraction. To perform these various functions, VGCCs are highly regulated. Inactivation comprises a critical mechanism controlling the entry of Ca2+ through these channels and constitutes an important means to regulate cellular excitability, shape action potentials, control intracellular Ca2+ levels, and contribute to long-term potentiation and depression. For CaV1 and CaV2 channel families, inactivation proceeds via two distinct processes. Voltage-dependent inactivation (VDI) reduces Ca2+ entry through the channel in response to sustained or repetitive depolarization, while Ca2+-dependent inactivation (CDI) occurs in response to elevations in intracellular Ca2+ levels. These processes are critical for physiological function and undergo exquisite fine-tuning through multiple mechanisms. Here, we review known determinants and modulatory features of these two critical forms of channel regulation and their role in normal physiology and pathophysiology.

Introduction

Voltage-gated calcium (Ca2+) channels (VGCCs) are crucial conduits for Ca2+ entry in both excitable and select non-excitable cells (Nowycky et al., 1985; Nilius et al., 1986; Pitt et al., 2021). They are integral to various physiological processes, including gene expression, hormone secretion, neurotransmitter release, synaptic function, and muscle contraction (Bean, 1989; Bers, 1991; Wu et al., 1999; Ma et al., 2011). Given the critical roles of VGCCs in both normal and pathological cellular functions, precise regulation and fine-tuning of their properties are essential (Lee et al., 1985). To this end, these channels employ two distinct forms of negative feedback regulation—voltage-dependent inactivation (VDI) and Ca2+-dependent inactivation (CDI). These regulatory processes are critical for the physiological function of the channel, enabling channels to modulate Ca2+ entry due to sustained activity, thus preventing the accumulation of excessive intracellular Ca2+ and contributing to dynamic physiological processes.

VGCCs are composed of a main α1 subunit and auxiliary subunits that can modulate channel properties, including β, α, and γ. The α1 subunit has four homologous domains, each containing six transmembrane α helices, designated S1–S6. The S1–S4 segments compose the voltage-sensing domain, with the S4 containing positive gating charges, which enable the channel to respond to changes in membrane voltage (Flucher, 2016). The S5 and S6 helices form the pore of the channel while the S5–S6 linker forms a loop containing the selectivity filter (Sather et al., 1994; Stephens et al., 2015). Lastly, S6 helices form the activation gate intracellularly.

There are three families of VGCCs based on their α1 subunits. These include CaV1 (CaV1.1–1.4), CaV2 (CaV2.1–2.3), and CaV3 (CaV3.1–3.3), also known as L- (CaV1), P/Q- (CaV2.1), N- (CaV2.2), R-type (CaV2.3), and T-type (CaV3) (Catterall et al., 2005). CaV1 and CaV2 channel families are often referred to as high voltage-activated channels; however, the term may be somewhat of a misnomer for CaV1.3 channels, which activate at more hyperpolarized voltages as compared with other L-type channels (Avery and Johnston, 1996; Xu and Lipscombe, 2001). CaV3 channels are part of the low voltage-activated Ca2+ channel subfamily and are activated at more hyperpolarized membrane potentials, have more rapid inactivation kinetics, and smaller unitary conductance as compared with CaV1 and CaV2 channels (Nowycky et al., 1985; Nilius et al., 1986; Zamponi et al., 2015; Nanou and Catterall, 2018). The inactivation of CaV3 channels appears to differ from that of CaV1 and CaV2 channel families in terms of structural elements, modulation by auxiliary subunits, and presence of CDI. We therefore focus this review on the CaV1 and CaV2 channel families.

CaV1.2 and CaV1.3 channels are widely expressed across excitable cells and are responsible for critical cellular functions such as excitation–contraction coupling, excitation-transcription coupling, synaptic regulation, and hormone release (Bers, 2002; Helton et al., 2005; Striessnig et al., 2006; Zhang et al., 2006). Conversely, CaV1.1 and CaV1.4 are more selectively expressed in skeletal muscle and retinal cells, respectively (Catterall et al., 2005). CaV2.1 channels are highly expressed in presynaptic terminals throughout the brain and are enriched in cerebellar Purkinje cells, where they play a critical role in balance and movement (Indriati et al., 2013). CaV2.2 channels are expressed throughout the nervous system and contribute to synaptic maturation, gene expression, neuronal migration and survival, and fast neurotransmitter release (Komuro and Rakic, 1992; Simms and Zamponi, 2014; Gorman et al., 2019). CaV2.3 channels are highly expressed in the central and peripheral nervous systems (Wormuth et al., 2016) and have been identified in the endocrine and cardiovascular systems (Lu et al., 2004). Channel regulation constitutes a major factor enabling the modulation of channel gating relevant to each physiological context. Here, we review the mechanisms and physiological implications of CDI and VDI in CaV1 and CaV2 channels.

Voltage-dependent inactivation

20 years after Hodgkin and Huxley postulated the existence of gating change (Hodgkin et al., 1952), two seminal papers provided experimental evidence for the movement of a voltage sensor within a membrane protein (Armstrong and Bezanilla, 1973; Schneider and Chandler, 1973). Schneider and Chandler were the first to measure the charge movement of CaV1.1 channels associated with excitation–contraction in skeletal muscle (Schneider and Chandler, 1973), a pivotal component in understanding how changes in membrane voltage led to conformational changes in the channel. This voltage-sensor movement was critical not only for channel activation but for VDI (Armstrong and Bezanilla, 1977), a critical feedback mechanism that can control the precise amount of Ca2+ influx through VGCCs (see Catacuzzeno et al. [2023] for a comprehensive review of the role of gating change in channel inactivation). Disruption of this process can lead to various pathological conditions, including cardiac arrhythmia, epilepsy, and chronic pain (Splawski et al., 2004; Gambeta et al., 2021; Herold et al., 2023; Hussey et al., 2023a).

Multiple mechanisms for VDI have been considered. The first involves a cytoplasmic segment of the channel that binds to and blocks the channel pore (Armstrong and Bezanilla, 1977; Bezanilla and Armstrong, 1977). This type of mechanism has been described as a “ball and chain” mechanism for the N-type inactivation of the K+ channels (Hoshi et al., 1990) and a “hinged lid” (IFM motif in the domain III–IV loop) for voltage-gated Na channels (West et al., 1992). However, structural evidence has demonstrated that for Na channels, the block of the intracellular gate does not occur via direct block by the IFM motif, but rather by insertion of the IFM into a cavity outside the S6 bundle. The effect of this is predicted to decouple the pore domain and the S4–S5 ring, causing the S6 helices to close (Huang et al., 2024). Interestingly, this “door wedge” model of inactivation shares many features of the hinged lid, making them difficult to distinguish experimentally (Fig. 1 A). The next mechanism involves a collapse of the selectivity filter or the P-loop in the S5–S6 linker as observed in the C-type inactivation (also known as P-type inactivation) of the K+ channels (Hoshi et al., 1990; Choi et al., 1991). Finally, VDI could potentially proceed from an allosteric mechanism in which the gating of the channel is modulated (Tadross et al., 2010).

The kinetics of VDI varies based on α1 subunit identity and channel-binding partners. VDI in CaV1.1 is kinetically slow compared with other CaV1 channels (Bannister and Beam, 2013), and the kinetics of VDI of CaV1.2 and CaV1.3 are faster than that of CaV1.4 (Koschak et al., 2003) and have been shown to display two kinetically distinct components (Ferreira et al., 2003). In CaV1 channels, VDI primarily occurs following the initial opening of the channel. Upon depolarization, the α1 subunit undergoes a conformational change, initiated by the movement of the VSD, and results in the opening of the intracellular activation gate, comprised of the S6 segments (Xie et al., 2005). With continued depolarization, the channel rapidly enters an inactivated state. Experimental evidence provides support for either a hinged lid or door wedge mechanism underlying the fast N-type-like inactivation seen in CaV1.2, CaV1.3, and CaV1.4 channels (Fig. 1 A) (Stotz et al., 2004). The “lid” or “wedge” in these channels appears to be comprised of the loop between homologous domains I and II, which has been proposed to interact with the distal S6 regions. A few lines of evidence support such a mechanism for CaV1 channels. First, VDI seen at the single-channel level is demonstrated as a complete loss of channel opening during sustained depolarization (Fig. 1 B) (Reuter et al., 1982; Yue et al., 1990), indicating complete closure or block of the channel. Next, multiple chimeric and point mutation studies demonstrate the importance of the S6 region in the initiation of VDI (Kraus et al., 1998; Stotz et al., 2000, 2004; Stotz and Zamponi, 2001a, 2001b; Splawski et al., 2004, 2005; Hoda et al., 2005; Hohaus et al., 2005; Raybaud et al., 2006; Barrett and Tsien, 2008; Dick et al., 2016; Bamgboye et al., 2022a). For example, the slower VDI of CaV1.2 can be sped up by replacing the S6 region of domain II with a homologous region from a more rapidly inactivating CaV2.3 channel (Stotz et al., 2000), while individual point mutations within the same region significantly modulate VDI (Stotz and Zamponi, 2001a). Overall, these experiments point to the importance of the distal S6 region in CaV1 VDI but do not distinguish between the hinged lid and the door wedge mechanisms. Next, mutations and chimeric studies aimed at the I–II loop of the channel result in similar alterations of VDI (Bourinet et al., 1999; Stotz et al., 2000, 2004; Bernatchez et al., 2001; Berrou et al., 2001; Stotz and Zamponi, 2001a, 2001b; Dafi et al., 2004). Interestingly, overexpression of the I–II loop region by itself was sufficient to increase VDI (Cens et al., 1999), indicating a role of the I–II loop in CaV1 VDI.

The I–II loop of VGCCs also contains an α-interaction domain (AID), the known binding site of the auxiliary β subunit (De Waard et al., 1995). As such, both the kinetics and magnitude of VDI can be modulated by the interaction with the β subunit, further supporting the role of the I–II loop in CaV1 inactivation. Transferring the AID segment from α1C to the α1E backbone was shown to slow VDI (Bernatchez et al., 2001), while mutations introduced within the AID of CaV1.2 (α1c) (Dafi et al., 2004), CaV1.3 (α1D) (Tadross et al., 2010), or CaV2.1 (α1A) led to significant alterations of VDI (Kraus et al., 1998). Furthermore, modulation of VDI is dependent on the β subunit variant. In particular, β2a is palmitoylated, such that the subunit is anchored to the membrane, thus restricting the mobility of the I–II loop and impairing VDI (Fig. 1 C) (Olcese et al., 1994; Herlitze et al., 1997; Stephens et al., 2000). On the other hand, β1 subunits which lack this palmitoylation allow the free movement of the I–II loop, permitting VDI. Interestingly, this regulatory process appears to be further fine-tuned in select channels. CaV1.3 has been proposed to contain a “shield” that prevents the complete closure of the hinged lid, leading to minimal VDI regardless of the type of β subunit. As evidence, mutagenesis of the shield in CaV1.3 channels renders them sensitive to β subunit isoforms (Tadross et al., 2010). However, these experiments could be explained by either a hinged lid or door wedge mechanism, where the shield mutation similarly obstructs the action of the I–II loop.

More recent structural studies have further demonstrated the complex nature of CaV1 VDI and call into question the hinged lid hypothesis. While each CaV1 structure has validated the distal S6 regions as the intracellular gate (Wu et al., 2015; Yao et al., 2022; Chen et al., 2023; Gao et al., 2023), they fail to show the I–II loop acting as a blocking particle. While this may point toward a door wedge mechanism, this feature has not yet been described. Thus, the detailed mechanism underlying the VDI of CaV1 channels remains to be fully elucidated.

In addition to fast N-type-like inactivation, CaV1 channels have also been shown to exhibit slower C-type-like inactivation (Catacuzzeno et al., 2023). For CaV1.1, this form of inactivation represents the dominant mechanism of channel inactivation and is associated with a collapse of the selectivity filter as described for C-type-like inactivation (Hoshi et al., 1990; Choi et al., 1991; Catacuzzeno et al., 2023). Likewise, CaV1.2 channels have been shown to undergo two kinetically distinct components of VDI (Ferreira et al., 2003). The slow component of CaV1.2 VDI may be explained by a similar C-type inactivation mechanism. In support of such a mechanism, the VDI of both CaV1.1. and CaV1.2 is accompanied by a characteristic negative shift in the voltage dependence of the gating charge (Brum et al., 1988; Pizarro et al., 1989; Ferreira et al., 2003; Villalba-Galea et al., 2008). Moreover, mutagenesis of the selectivity filter of CaV1.3 channels demonstrates a potential role for the selectivity filter in modulating VDI (Del Rivero Morfin et al., 2024). Mutation of a highly conserved domain IV tryptophan within the selectivity filter significantly enhances VDI in a β subunit–independent manner. Moreover, this form of VDI exhibited a distinct U-shaped dependence on voltage, which was insensitive to the identity of the charge carrier. Thus, it appears that CaV1 channels are capable of undergoing a distinct form of VDI resulting from a mechanism of selectivity filter collapse, pointing to more than one VDI mechanism at play. Moreover, the VDI of CaV1 channels may be further modulated by interaction with other proteins, including the adaptor protein STAC3, which has been shown to reduce the VDI of CaV1.1 (Tuinte et al., 2022) and CaV1.2 (Wong King Yuen et al., 2017) channels.

Compared with CaV1 channels, CaV2 channels have much faster kinetics (onset/recovery) of VDI (Stotz et al., 2000). Both CaV2.3 and CaV2.1 channels possess fast N-type-like inactivation, although CaV2.3 inactivation is shown to be faster than CaV2.1 (Zhang et al., 1994; Herlitze et al., 1997). CaV2.2 possesses both N-type-like and relatively slower inactivation similar to C-type inactivation (McDavid and Currie, 2006; Zhu et al., 2015). Similar to CaV1, the magnitude and kinetics of VDI in CaV2 are also modulated by types of β subunits (Buraei and Yang, 2013). However, unlike CaV1, inactivation of CaV2.2 and CaV2.1 can be further modulated by G protein βγ subunits (Herlitze et al., 1997; McDavid and Currie, 2006). Additionally, CaV2 channels also display a U-shaped dependence on voltage, independent of the ionic current (Patil et al., 1998). For these channels, VDI has been shown to proceed preferentially from an intermediate closed-state channel conformation (Patil et al., 1998; Jones et al., 1999). Structural evidence for CaV2.2 channels indicates that the channels contain a W-helix within the domain II–III linker, which locks the intracellular gate of the channel in the closed configuration (Dong et al., 2021). Within this closed-inactivated state, the side chain of residue W768 of the II–III linker extends into and blocks the intracellular gate, while multiple residues within the W-helix form hydrophobic interactions with residues residing in the distal S6 regions. Thus, the closed-state inactivation of CaV2 channels partially mirrors a hinged lid mechanism.

Ca2+-dependent inactivation

Modulation of CaV channels by intracellular Ca2+ was first reported in Paramecium by Brehm and Eckert (1978). They noticed that the inward Ca2+ current “relaxed within 10 ms” of initiation and that this inactivation disappeared when extracellular Ca2+ was replaced by strontium or barium. Since this remarkable discovery, a tremendous amount of cumulative work further sheds light on the Ca2+-dependent regulation of VGCCs, which takes the form of either CDI or Ca2+-dependent facilitation (CDF), where the channel open probability decreases or increases in response to a rise in cytosolic Ca2+, respectively.

Although there are several channel regions with Ca2+-binding capacity, including the EF hand-like region in the C-terminus of the α1 subunit itself, calmodulin (CaM), a bilobal Ca2+-binding protein, is shown to be the central driver of both CDI and CDF in CaV1 and CaV2 channels (Lee et al., 1999; Zühlke et al., 1999; Lee et al., 2000; DeMaria et al., 2001). As evidence, overexpression of a CaM-binding inhibitor peptide and deletion of the CaM-binding region in the α1A subunit of CaV2.1 channels render the channel unresponsive to changes in intracellular Ca2+ concentrations (Lee et al., 1999, 2000). Moreover, mutagenesis of the four EF hands, the Ca2+-binding motifs, across the N- and C-lobes of CaM (CaM1234) eliminates CDI in L-type channels and both CDI and CDF of CaV2.1 (Peterson et al., 1999; Zühlke et al., 1999; DeMaria et al., 2001). Of note, this striking dominant negative effect of CaM1234 in eliminating Ca2+-dependent regulation hints at the fact that CaM can interact with the α1 subunit in a Ca2+-independent manner. ApoCaM (Ca2+-free CaM) pre-associates with the IQ domain of the α1 subunit, as seen via FRET live-cell imaging (Erickson et al., 2001, 2003) and dansyl–CaM binding (Zühlke et al., 1999; Pitt et al., 2001). This pre-associated apoCaM acts as a resident Ca2+ sensor for the VGCCs. At resting or low levels of cytosolic Ca2+, the C-lobe of apoCaM engages the IQ domain in the C-terminus of the α1 subunit, while the N-lobe appears to associate weakly with dual vestigial EF hands upstream from the IQ domain (Fig. 2 A) (Erickson et al., 2001, 2003; Pitt et al., 2001; Ben Johny et al., 2013). Upon the rise of intracellular Ca2+, both lobes of CaM are calcified and mobilized to various segments of the N- and C-termini of the α1 subunit, initiating CDI. The IQ region has been identified as a primary locus for calcified C-lobe CaM binding (Lee et al., 2003; Chaudhuri et al., 2004; Van Petegem et al., 2005), possibly forming a tripartite complex with the two vestigial EF hands within the proximal C-terminus of the channel (Bazzazi et al., 2013; Ben Johny et al., 2013). The locus for the calcified N-lobe, on the other hand, is less well-defined. While the calcified N-lobe of CaM is capable of binding the IQ domain (Van Petegem et al., 2005; Mori et al., 2008; Minor and Findeisen, 2010), mutagenesis studies indicate that this interaction may not be functionally relevant for CDI within the holo-channel (Mori et al., 2008). For CaV1.2 and CaV1.3 channels, the calcified N-lobe of CaM has been shown to bind to an N-terminal spatial Ca2+-transforming element (NSCaTE) within the amino terminus of the channel (Ivanina et al., 2000; Dick et al., 2008; Tadross et al., 2008). However, a similar motif has not been identified for other channels.

Unlike VDI, where channel opening is completely blocked, CDI proceeds via an allosteric mechanism, whereby Ca2+ binding to CaM results in a transition of the channel from a high open probability mode (mode 1) to a lower open probability mode (mode Ca) (Fig. 2 B). At the single-channel level, this allosteric mechanism can be visualized as persistence in channel openings through the duration of a depolarization (Fig. 2 C) with a reduced open probability as compared with the initial mode 1 openings seen at the start of the depolarization in Ca2+ (see bottom averaged trace) or throughout the step in the absence of Ca2+ (Fig. 1 B) (Yue et al., 1990; Imredy and Yue, 1994). The structural determinants of this reduced channel opening remain largely unknown; however, domain II of the selectivity filter has been identified as a critical element, implicating the collapse of the selectivity filter as a potential endpoint for CDI (Abderemane-Ali et al., 2019). Thus, VGCCs may utilize a mechanism similar to the C-type inactivation of K+ channels for CDI.

CaM regulation of VGCCs displays a remarkable bipartition of regulation, such that each lobe of CaM can impart a distinct and independent form of channel regulation. The initial observation of this functional bipartition again emerged from observations in Paramecium, where mutations to residues within the N-lobe of CaM rendered Paramecium under reactive to stimuli, while mutations to residues in the C-lobe of CaM led to over reactive Paramecium, reflecting loss of either a Ca2+-dependent Na+ current or a Ca2+-dependent K+ current, respectively (Kink et al., 1990). Invaluable tools with targeted ablation of Ca2+-binding to the EF hands of N- or C-lobe CaM were developed to further dissect the Ca2+-dependent regulatory processes of ion channels (Xia et al., 1998). Mutagenesis of the two EF hands in the N-lobe (CaM12) enables evaluation of C-lobe Ca2+ regulation, while mutagenesis of the EF hands in the C-lobe (CaM34) provides evaluation of N-lobe CaM regulation (Liang et al., 2003). Overexpression of either CaM12 or CaM34 results in robust CDI in CaV1.3, visualized as the stronger decay of the Ca2+ current as compared with when Ba2+ is used as the charge carrier (Fig. 3 A), demonstrating the ability of either lobe to impart CDI. However, the kinetics of CDI differ significantly between the two lobes, demonstrating that each lobe is capable of imparting distinct regulatory features. For CaV2.1 channels, known to undergo both CDI and CDF, this bipartition of function is even more dramatic, with the N-lobe of CaM supporting CDI, while the C-lobe is responsible for CDF (Zühlke et al., 1999; DeMaria et al., 2001). On the other hand, CaV2.2 and CaV2.3 channels appear to only exhibit CDI due to N-lobe calcification (Liang et al., 2003). Thus, each lobe of CaM can impart an independent form of Ca2+ regulation, differing across channel subtypes (Fig. 3).

In addition to functional bipartition, CaM also exhibits spatial selectivity in Ca2+ sensing. This phenomenon is unveiled by the utilization of various types of Ca2+ buffers. In the presence of low or physiological Ca2+ buffering, Ca2+ entry consists of two major components. The local Ca2+ signal is comprised of rapid spikes of Ca2+ resulting from individual channel openings and closings (Fig. 3 B). A second, smaller global Ca2+ signal arises from the accumulation of cytosolic Ca2+ due to the aggregate opening of multiple channels and Ca2+ sources throughout the cell. Strong buffers may be used to restrict Ca2+ to the nanodomain, preventing the global accumulation of Ca2+ and enabling evaluation of the functional effect of local Ca2+ signaling (Stern, 1992). Interestingly, each lobe of CaM can distinguish between these two kinetically distinct Ca2+ signals. The C-lobe of CaM invariably responds to local Ca2+, such that regulation persists even in the presence of high intracellular Ca2+ buffering (Fig. 3 C). On the other hand, the N-lobe of CaM often requires a global elevation in Ca2+, which is only present under low intracellular Ca2+ buffering (DeMaria et al., 2001; Lee et al., 2003; Dick et al., 2008; Tadross et al., 2008). The mechanism by which this spatial Ca2+ selectivity comes about stems from the kinetics of Ca2+ binding to each lobe of CaM, where the C-lobe exhibits relatively slow Ca2+ (un)binding kinetics (slow CaM mechanism) ensuring a response to the large local Ca2+ spikes (Tadross et al., 2008). The N-lobe, on the other hand, can rapidly (un)bind Ca2+ (slow-quick-slow mechanism), enabling the lobe to respond to either the slow global Ca2+ signal or the rapid Ca2+ spikes of the local signal, dependent on the affinity of the calcified lobe for the channel. Thus, N-lobe spatial selectivity can be tuned for each channel subtype (Fig. 3 C). While the majority of VGCCs exhibit global Ca2+ selectivity for N-lobe CaM-mediated regulation, CaV1.2 and CaV1.3 channels contain the NSCaTE Ca2+/CaM-binding motif, which binds to the N-lobe of CaM, modifying the spatial selectivity of CaV1.2 and CaV1.3 channels from a global to a local selectivity (Dick et al., 2008; Tadross et al., 2008).

Biological fine-tuning of CDI

As a critical feedback mechanism, CDI can be modified by a variety of mechanisms, enabling fine-tuning of channel regulation suited to distinct biological needs. First, alternative splicing of the genes encoding the α1 subunits may modify the elements responsible for CDI within select channel variants, resulting in significant changes in CDI. The distal C-terminus of CaV1.3 and CaV1.4 channels are known to contain a region known as an inhibitor of CDI (ICDI) or C-terminal modulatory (CTM) domain (Singh et al., 2006; Wahl-Schott et al., 2006; Sang et al., 2021) that can compete with apoCaM for binding to the IQ domain. This interaction results in a decreased open probability of the channel (Adams et al., 2014) and loss of CDI and can be modulated by cytosolic expression levels of CaM (Liu et al., 2010). The inclusion of ICDI/CTM is highly regulated by alternative splicing. In CaV1.3 channels, exon 42 gives rise to a long form of the channel containing ICDI/CTM, which exhibits minimal CDI. The inclusion of alternative exon 42A or utilization of an alternative 3′ splice-acceptor site in exon 43 produces a truncated channel bereft of the motif, thus restoring CDI (Singh et al., 2008; Bock et al., 2011). Additionally, the splicing of exons 44 and 48 modulates CDI by altering ICDI/CTM interaction with the IQ domain (Tan et al., 2011). Similar features were identified in CaV1.4, which does not display CDI due to the presence of ICDI/CTM in the full-length channel, but exhibits significant CDI upon either deletion of exon 47 (Williams et al., 2018) or inclusion of exon 43 containing a premature stop codon (Tan et al., 2012). Moreover, phosphorylation of this domain in CaV1.4 can further modulate the interaction, enabling dynamic regulation of CDI within select cell types (Sang et al., 2016). Thus, modification of the ICDI/CTM interaction with the IQ domain represents a highly tunable CDI modulatory mechanism in CaV1.3 and CaV1.4 channels.

Another example of alternative splicing as a method to diversify channel properties is found in CaV2.1. Splicing at the EF hands in (exon 37) and C-terminus distal to the CBD domain (exon 47) modulates the magnitude of CDF without altering CDI (Chaudhuri et al., 2004). The distribution of these splice variants is heterogeneous in the human brain, showing both spatial variation across brain regions and temporal variation across stages of brain development. This added layer of complexity likely allows cells to fine-tune the properties of CaV2.1 channels to best suit their functional requirement across time and space.

In addition to splice variation, RNA editing represents another mechanism that enables the modulation of CDI. For CaV1.3, the IQ motif is subject to significant RNA editing, resulting in channels with altered affinity between the channel and apoCaM (Bazzazi et al., 2013; Huang et al., 2012). By modulating the interaction between the channel IQ region and ICDI/CTM, RNA editing tunes the magnitude of CDI in CaV1.3. Notably, the pattern of RNA editing varies across different regions of the brain, rendering a landscape of CaV1.3 channels with varying degrees of CDI that best serves their functional requirements (Huang et al., 2012).

Further fine-tuning of CDI can occur through the interaction of the channel with multiple proteins. The CaM-like calcium-binding protein (CaBP) has been shown to bind to CaV1.2 and CaV1.3, both competitively against CaM and allosterically (at a binding site distinct from the IQ region), resulting in significant changes in the magnitude of CDI (Yang et al., 2006; Findeisen and Minor, 2010; Yang et al., 2014; Hardie and Lee, 2016). CaBP4 is highly expressed in inner hair cells and impedes CDI of CaV1.3 (Yang et al., 2006, 2014), a modulation necessary for continued perception of sustained auditory stimuli (Lewis and Hudspeth, 1983). CaBP1, found primarily in neurons, inhibits CDI and unveils CDF of CaV1.2 (Zhou et al., 2005; Findeisen and Minor, 2010), further demonstrating the complexity of channel regulation in the brain. Additionally, SH3 and cysteine-rich domain (STAC) can bind to and modulate the gating of VGCCs (Polster et al., 2015, 2018; Campiglio et al., 2018; Niu et al., 2018b; Flucher and Campiglio, 2019). STAC1 and STAC2 are expressed primarily in the brain (Suzuki et al., 1996; Nelson et al., 2013), peripheral nervous system (Legha et al., 2010), retina (Wilhelm et al., 2014), and inner ear (Cai et al., 2015), while STAC3 is expressed in the skeletal muscle (Nelson et al., 2013). All three STAC isoforms are shown to impede CDI of CaV1.2 and CaV1.3 without alteration of CDI in CaV2.1 channels (Niu et al., 2018a; Polster et al., 2018). These “third-party” binding partners add yet another layer of sophistication to the Ca2+-dependent regulatory processes of VGCCs.

CDI and VDI in physiology

VGCCs are present in all excitable cells such that channel inactivation profiles have the potential to impact a myriad of physiological functions. One of the most well-recognized roles for VGCC inactivation occurs in the heart, where CaV1.2 plays a critical role in shaping action potential (AP) morphology. The channel opens primarily during the plateau phase of the AP, and inactivation occurs during the sustained depolarization, reducing the depolarizing effect of the channel. The importance of this inactivation process has long been recognized, yet it is not always trivial to identify the relative impact of VDI versus CDI. In fact, computational models of the cardiac AP demonstrate that either or both VDI and CDI can support the inactivation required to produce a physiological AP (Livshitz and Rudy, 2007; Morotti et al., 2012). Reduction of CDI via overexpression of either CaM1234 or CaM34 in adult guinea pig ventricular myocytes results in a profound prolongation of the AP (Fig. 4 A) (Alseikhan et al., 2002), demonstrating the importance of CaM in controlling the AP duration and implicating CDI as a critical regulator of cardiac AP morphology. Yet, the conclusion is complicated by the myriad of CaM effectors in the heart and does not address the potential impact of VDI. To address this, Morales et al. (2019) utilized a combination of CaM mutants and overexpressed β subunit variants to demonstrate distinct effects of CDI and VDI on cardiomyocyte function, where disruption of either process modulated the AP at baseline, with a larger effect due to CDI disruption. In addition, under conditions of β-adrenergic stimulation, CDI was shown to be of paramount importance as compared with VDI. Thus, both processes contribute significantly to the AP morphology, yet CDI appears to exert a greater impact.

In the brain, VGCCs play important roles in synaptic transmission, excitation–transcription coupling, and long-term potentiation. Altered Ca2+ entry through multiple VGCCs can disrupt numerous neurological functions, yet the specific role of inactivation in these processes is often difficult to determine. While the rapid time course of a neuronal AP reduces the importance of inactivation in shaping a single AP, as seen in cardiac myocytes, trains or bursts of APs may be modulated by VGCC inactivation properties. For CaV2.1, CDI and CDF have been shown to play an important role in synaptic plasticity (Cuttle et al., 1998; Tsujimoto et al., 2002; Xu and Wu, 2005; Mochida et al., 2008; Adams et al., 2010) and have been implicated in long-term potentiation and memory (Nanou et al., 2016). Moreover, numerous processes in the brain have been shown to modulate the extent of VGCC inactivation in a tissue-specific manner. RNA editing of the CaV1.3 channel has been shown to reduce CDI and occurs in as many as half the transcripts in the brain (Huang et al., 2012) and is controlled in a neuron-specific manner by ADAR (Huang et al., 2018). This editing is thought to play a role in the rhythmicity of the suprachiasmatic nucleus (Huang et al., 2012) and in synaptic plasticity, learning, and memory (Zhai et al., 2022). In addition, splice variation has been shown to modify the inactivation of multiple VGCCs, demonstrating a role for inactivation modulation in neurophysiology. CaV2.1 channels contain alternative exons within the C-terminus of the channel known to significantly modulate CDI and CDF (Soong et al., 2002). These variants are differentially expressed (Bunda and Andrade, 2022) and are known to modulate the pathogenesis of select CaV2.1 variants (Adams et al., 2009, 2010) and factor into protecting the brain from disease (Aikawa et al., 2017), demonstrating the physiological importance of CDI and CDF. For CaV1.3 and CaV1.4, the presence of ICDI/CTM within the long channel variant significantly suppresses CDI (Scharinger et al., 2015; Williams et al., 2018), enabling tissue-specific suppression of inactivation in the retinal ganglion, chromaffin, and inner hair cells, and modulation of CDI across different tissues and in development (Lieb et al., 2012; Scharinger et al., 2015; Williams et al., 2018).

In addition, multiple subunits are known to modify VGCC regulation, demonstrating the importance of tuning VDI and CDI in distinct contexts. For example, the α2δ subunit has been shown to accelerate channel inactivation, and co-expression of α2δ with CaV2.1 and CaV2.2 resulted in increased coupling to exocytosis and enhanced synaptic transmission (Hoppa et al., 2012). On the other hand, the β2A subunit strongly diminishes the VDI of most VGCCs as compared with other β subunits and can have significant effects on synaptic function (Wemhöner et al., 2015). Physiologically, this phenomenon has been associated with asynchronous transmission resulting from prolonged openings of CaV2.1 at the synapse (Müller et al., 2010). RIM proteins modulate channels both through interaction with the β subunit and G-protein receptors, causing a decrease in VDI of CaV2.2 (Weiss et al., 2011). Likewise, RIM2α was shown to slow the inactivation of CaV1.3 in inner hair cells (Gebhart et al., 2010). Such specialized modulation of inactivation appears to be of particular importance in the auditory and visual systems, where slow inactivation of CaV1.3 and CaV1.4 may be critical to enable the graded responses of these synapses (McRory et al., 2004; Inagaki and Lee, 2013). For CaV1.3 in cochlear inner hair cells, VDI is suppressed in a splice-dependent manner by a combination of β subunit identity and association of the channel with presynaptic scaffolding proteins, including RIM2α and RBP2 (Ortner et al., 2020b). CDI in these cells is likewise suppressed by the presence of CaBP4 (Yang et al., 2006), as well as the expression of the long ICDI/CTM-containing channel variant (Scharinger et al., 2015). Overall, numerous mechanisms impart exquisite control over inactivation across different tissues and throughout development, demonstrating the importance of CDI and VDI in physiology.

Role of channel inactivation in disease pathogenesis

Numerous genetic mutations involved in the disruption of channel inactivation have been associated with severe clinical manifestations. Timothy syndrome (TS), also known as long-QT syndrome type 8 (LQT8), is a multisystem disorder often including profound long-QT syndrome (LQTS) and neurodevelopmental disorders (NDD), resulting from a genetic mutation within CaV1.2. The canonical TS mutation, G406R, causes a profound reduction in both CDI and VDI (Fig. 4 B) (Splawski et al., 2004, 2005; Barrett and Tsien, 2008; Dick et al., 2016). Like G406R, many TS mutations impact CDI, VDI, and activation (Splawski et al., 2004, 2005; Barrett and Tsien, 2008; Boczek et al., 2015; Wemhöner et al., 2015; Dick et al., 2016; Landstrom et al., 2016; Bamgboye et al., 2022a), making it difficult to determine the relative impact of each form of channel regulation. However, selected mutations associated with LQTS have been shown to exhibit more selective effects on CDI or VDI (Bamgboye et al., 2022a), demonstrating that each form of channel regulation can contribute significantly to the cardiac features of the disorder. Importantly, even a modest alteration of CDI (Bamgboye et al., 2022a) may correlate with LQTS and sudden cardiac death (Wemhöner et al., 2015). Conversely, one mutation (L762F), which appears to have a more selective effect on VDI (Bamgboye et al., 2022a), correlated with a somewhat less severe presentation of LQTS across multiple members of a single family (Landstrom et al., 2016). However, it is difficult to draw direct conclusions due to limited known occurrences of the mutations and the existence of other cardiac variants within the family members harboring the L762F mutation (Landstrom et al., 2016). Nonetheless, the pattern of biophysical effects (Bamgboye et al., 2022a) thus far appears to point to more severe cardiac phenotypes in patients harboring mutations involving a loss of CDI. Consistent with the idea that CDI plays a dominant role in determining the cardiac AP, genetic mutations in CaM (calmodulinopathies), which disrupt Ca2+ binding to CaM and thus severely diminish CDI of CaV1.2, also cause a profound prolongation of the cardiac AP and severe LQTS (Limpitikul et al., 2014; George, 2015; Jensen et al., 2018; Crotti et al., 2019; Nyegaard and Overgaard, 2019). Thus, in the heart, disruptions of either VDI or CDI can dramatically prolong the AP and cause severe pathologies, although decreases in CDI appear likely to produce more significant phenotypes.

In the brain, TS mutations have been linked to NDD, pointing to a causative role for overt increases in Ca2+ entry (Splawski et al., 2004; Limpitikul et al., 2016; Pinggera and Striessnig, 2016; Pinggera et al., 2017). However, unlike in the heart, a severe NDD phenotype associated with many TS patients (often presenting as autism spectrum disorder) appears to correlate more consistently with hyperpolarizing shifts in channel activation rather than inactivation (Marcantoni et al., 2020; Bamgboye et al., 2022a). Similar mutations within the CaV1.3 channel appear to follow an analogous pattern, where mutations alter multiple channel biophysical properties, but a severe NDD phenotype appears to correlate largely with activation changes (Pinggera et al., 2015; Ortner et al., 2023; Dannenberg et al., 2024). Nonetheless, this result represents only the severe neurodevelopmental deficits historically described in TS. As the recognition of patients harboring CaV1.2 mutations has increased, so too has the variety of reported symptoms and phenotypes, with a particular increase in recognized neurological symptoms (Gillis et al., 2012; Bozarth et al., 2018; Rodan et al., 2021; Levy et al., 2023; Timothy et al., 2024). Moreover, patients harboring mutations in CaM (calmodulinopathies) often exhibit neurological features, including seizures and neurodevelopmental delay (Crotti et al., 2013; Retterer et al., 2016; Takahashi et al., 2016; Crotti et al., 2019; Hussey et al., 2023b). Thus, it is likely that deficits in CDI play a critical role in neuropathogenesis.

While it is difficult to rule out hypoxic brain injury secondary to cardiac arrest in cases involving CaV1.2 mutations, multiple neurological phenotypes have been associated with inactivation deficits in VGCCs primarily expressed in the brain. Inactivation changes have been identified in multiple CaV1.4 mutations, resulting in congenital stationary night blindness type 2 (Hemara-Wahanui et al., 2005; Hoda et al., 2005, 2006). While many of these mutations do not represent selective effects on inactivation, truncation mutations that remove the CTM/ICDI region from the channel cause a predominant increase in CDI and congenital stationary night blindness type 2, demonstrating the importance of sustained (non-inactivating) current to support the tonic release of neurotransmitters at sensory cell ribbon synapses (Singh et al., 2006; Wahl-Schott et al., 2006; Griessmeier et al., 2009). Mutations in CaV2.1 associated with episodic ataxia type 2 have been shown to alter multiple channel properties, including modulation of inactivation (Spacey et al., 2004; Cuenca-León et al., 2009). In many cases, the distinction between CDI and VDI effects was not made, although the use of Ba2+ implicates increased VDI in the pathogenesis of select mutations (Spacey et al., 2004). Likewise, CaV2.1 mutations associated with familial hemiplegic migraine type 1 (FHM1) have also been shown to alter inactivation, in some cases increasing inactivation (Serra et al., 2009). Kinetic changes in VDI due to the S218L mutation in CaV2.1 have been implicated in the pathogenic mechanism underlying cortical spreading depression in FHM1 (Tottene et al., 2005). Alternative splicing of CaV2.1 channels, which alters multiple channel properties, including recovery from inactivation, modulates the pathogenesis of FHM1 (Adams et al., 2009). Likewise, mutations in CaV2.3, which are causative of developmental and epileptic encephalopathies have been identified within the distal S6 regions and shown to modulate channel activation and inactivation (Helbig et al., 2018; Sampedro-Castañeda et al., 2023). In addition, amyotrophic lateral sclerosis has been associated with a reduction in RNA editing of the frontal cortex resulting in alterations in CaV1.2. Karagianni et al. (2024), implicating altered CDI of CaV1.2 in the pathogenesis of the disease. Finally, calmodulinopathy mutations not only reduce CDI in CaV1.2 channels, but decrease CDI in CaV1.3 and diminish CDF in CaV2.1, potentially contributing to the neurological effects identified in some patients (Hussey et al., 2024, Preprint). Thus, the pathogenic effects of disrupted VGCC inactivation mechanisms appear to be a recurrent theme, with the identification of the specific impact of each form of channel regulation within different tissues representing an important question still to be resolved.

Implications for therapeutics

Many Ca2+ channel blockers (CCBs) are known to be state dependent, such that the drug interacts preferentially with the inactivated channel. The state-dependent block of CaV1 channels by multiple CCBs has been well characterized (Kanaya et al., 1983; Hering et al., 1997; Berjukow et al., 2000). CCBs are divided into three main classes: dihydropyridines, phenylalkylamines, and benzothiazepines. Dihydropyridines are FDA-approved CaV1 channel blockers, often used in the treatment of hypertension. These drugs bind to a hydrophobic pocket at the interface between domains III and IV of the pore domain, accessible via fenestrations, which are state dependent (Zhao et al., 2019; Gao and Yan, 2021). As a result, conformational changes modulate access of the drug to the binding site, resulting in an enhanced block of the inactivated state of the channel. Phenylalkylaminess and benzothiazepines bind to an overlapping site within the central cavity of the channel (Zhao et al., 2019) and also show preferential binding to the inactivated state of the channel. This state-dependent block of CCBs provides a use-dependent feature, where sustained or repeated depolarization increases channel block (Hering et al., 1997). Such state-dependent block can also be observed in CaV2 channels. Multiple state-dependent inhibitors of CaV2.2 have been identified as potential analgesics (Abbadie et al., 2010; Pajouhesh et al., 2010; Swensen et al., 2012), with a block of the channel biased toward open and inactivated states.

This inactivation-dependent block can further modulate the efficacy of these drugs in the context of inactivation-modulating channel mutations. Genetic mutations in VGCCs which reduce inactivation result in significant increases in Ca2+ current. Although CCBs appear to present a useful therapeutic option, multiple reports indicate a lack of efficacy in these patients. In particular, verapamil has been tried in the treatment of both TS and calmodulinopathies but failed to reduce the QT interval (Jacobs et al., 2006; Shah et al., 2012; Crotti et al., 2013; Webster et al., 2017). This lack of efficacy may be attributed to the state-dependent properties of CCBs, including verapamil (Johnson et al., 1996; Catterall, 2000), where the drug binds preferentially to the inactivated state of the channel. Thus, the drug does not inhibit the mutant channels lacking inactivation to the same degree as wild-type channels (Sheng et al., 2012; Bamgboye et al., 2022b), failing to rescue the prolonged AP produced by the TS mutation (Fig. 4 C). While this state-dependent effect may be a disadvantage for mutations causing deficits in inactivation, it could also represent an opportunity for channels where inactivation is enhanced, resulting in increased CCB efficacy on the mutant channel as has been shown for select CaV1.3 mutations (Hofer et al., 2020; Ortner et al., 2020a). Thus, channel inactivation stands as a critical mechanism in both normal physiology and pathophysiology, capable of impacting the progression of disease and options available for treatment. However, the role of inactivation in the pathogenesis of multiple disorders also points to a potential therapeutic target with significant advantages compared with a complete block of the channel. For example, roscovitine has been shown to enhance the VDI of CaV1.2. (Yarotskyy and Elmslie, 2007; Yarotskyy et al., 2010), making it a promising candidate for the treatment of LQTS and TS (Yarotskyy et al., 2009; Karagueuzian et al., 2017; Song et al., 2017; Angelini et al., 2021), potentially bypassing the limitations of CCBs.

Eduardo Ríos served as editor.

We thank all members of Ivy Dick’s lab at the University of Maryland School of Medicine for their support. We also thank Dr. Manu Ben-Johny from Columbia University for insightful comments and discussion.

This project was supported by a National Institutes of Health/National Heart, Lung, and Blood Institute grant (1R01HL149926).

Author contributions: W.B. Limpitikul: writing—original draft, review, and editing. I.E. Dick: conceptualization, funding acquisition, visualization, and writing—original draft, review, and editing.

Abbadie
,
C.
,
O.B.
McManus
,
S.Y.
Sun
,
R.M.
Bugianesi
,
G.
Dai
,
R.J.
Haedo
,
J.B.
Herrington
,
G.J.
Kaczorowski
,
M.M.
Smith
,
A.M.
Swensen
, et al
.
2010
.
Analgesic effects of a substituted N-triazole oxindole (TROX-1), a state-dependent, voltage-gated calcium channel 2 blocker
.
J. Pharmacol. Exp. Ther.
334
:
545
555
.
Abderemane-Ali
,
F.
,
F.
Findeisen
,
N.D.
Rossen
, and
D.L.
Minor
Jr.
2019
.
A selectivity filter gate controls voltage-gated calcium channel calcium-dependent inactivation
.
Neuron
.
101
:
1134
1149.e3
.
Adams
,
P.J.
,
E.
Garcia
,
L.S.
David
,
K.J.
Mulatz
,
S.D.
Spacey
, and
T.P.
Snutch
.
2009
.
Ca(V)2.1 P/Q-type calcium channel alternative splicing affects the functional impact of familial hemiplegic migraine mutations: Implications for calcium channelopathies
.
Channels
.
3
:
110
121
.
Adams
,
P.J.
,
R.L.
Rungta
,
E.
Garcia
,
A.M.
van den Maagdenberg
,
B.A.
MacVicar
, and
T.P.
Snutch
.
2010
.
Contribution of calcium-dependent facilitation to synaptic plasticity revealed by migraine mutations in the P/Q-type calcium channel
.
Proc. Natl. Acad. Sci. USA
.
107
:
18694
18699
.
Adams
,
P.J.
,
M.
Ben-Johny
,
I.E.
Dick
,
T.
Inoue
, and
D.T.
Yue
.
2014
.
Apocalmodulin itself promotes ion channel opening and Ca(2+) regulation
.
Cell
.
159
:
608
622
.
Aikawa
,
T.
,
T.
Watanabe
,
T.
Miyazaki
,
T.
Mikuni
,
M.
Wakamori
,
M.
Sakurai
,
H.
Aizawa
,
N.
Ishizu
,
M.
Watanabe
,
M.
Kano
, et al
.
2017
.
Alternative splicing in the C-terminal tail of Cav2.1 is essential for preventing a neurological disease in mice
.
Hum. Mol. Genet.
26
:
3094
3104
.
Alseikhan
,
B.A.
,
C.D.
DeMaria
,
H.M.
Colecraft
, and
D.T.
Yue
.
2002
.
Engineered calmodulins reveal the unexpected eminence of Ca2+ channel inactivation in controlling heart excitation
.
Proc. Natl. Acad. Sci. USA
.
99
:
17185
17190
.
Angelini
,
M.
,
A.
Pezhouman
,
N.
Savalli
,
M.G.
Chang
,
F.
Steccanella
,
K.
Scranton
,
G.
Calmettes
,
M.
Ottolia
,
A.
Pantazis
,
H.S.
Karagueuzian
, et al
.
2021
.
Suppression of ventricular arrhythmias by targeting late L-type Ca2+ current
.
J. Gen. Physiol.
153
:e202012584.
Armstrong
,
C.M.
, and
F.
Bezanilla
.
1973
.
Currents related to movement of the gating particles of the sodium channels
.
Nature
.
242
:
459
461
.
Armstrong
,
C.M.
, and
F.
Bezanilla
.
1977
.
Inactivation of the sodium channel. II. Gating current experiments
.
J. Gen. Physiol.
70
:
567
590
.
Avery
,
R.B.
, and
D.
Johnston
.
1996
.
Multiple channel types contribute to the low-voltage-activated calcium current in hippocampal CA3 pyramidal neurons
.
J. Neurosci.
16
:
5567
5582
.
Bannister
,
R.A.
, and
K.G.
Beam
.
2013
.
Ca(V)1.1: The atypical prototypical voltage-gated Ca²⁺ channel
.
Biochim. Biophys. Acta
.
1828
:
1587
1597
.
Bamgboye
,
M.A.
,
K.G.
Herold
,
D.C.O.
Vieira
,
M.K.
Traficante
,
P.J.
Rogers
,
M.
Ben-Johny
, and
I.E.
Dick
.
2022a
.
CaV1.2 channelopathic mutations evoke diverse pathophysiological mechanisms
.
J. Gen. Physiol.
154
:e202213209.
Bamgboye
,
M.A.
,
M.K.
Traficante
,
J.
Owoyemi
,
D.
DiSilvestre
,
D.C.O.
Vieira
, and
I.E.
Dick
.
2022b
.
Impaired CaV1.2 inactivation reduces the efficacy of calcium channel blockers in the treatment of LQT8
.
J. Mol. Cell. Cardiol.
173
:
92
100
.
Barrett
,
C.F.
, and
R.W.
Tsien
.
2008
.
The Timothy syndrome mutation differentially affects voltage- and calcium-dependent inactivation of CaV1.2 L-type calcium channels
.
Proc. Natl. Acad. Sci. USA
.
105
:
2157
2162
.
Bazzazi
,
H.
,
M.
Ben Johny
,
P.J.
Adams
,
T.W.
Soong
, and
D.T.
Yue
.
2013
.
Continuously tunable Ca(2+) regulation of RNA-edited CaV1.3 channels
.
Cell Rep.
5
:
367
377
.
Bean
,
B.P.
1989
.
Neurotransmitter inhibition of neuronal calcium currents by changes in channel voltage dependence
.
Nature
.
340
:
153
156
.
Ben Johny
,
M.
,
P.S.
Yang
,
H.
Bazzazi
, and
D.T.
Yue
.
2013
.
Dynamic switching of calmodulin interactions underlies Ca2+ regulation of CaV1.3 channels
.
Nat. Commun.
4
:
1717
.
Berjukow
,
S.
,
R.
Marksteiner
,
F.
Gapp
,
M.J.
Sinnegger
, and
S.
Hering
.
2000
.
Molecular mechanism of calcium channel block by isradipine. Role of a drug-induced inactivated channel conformation
.
J. Biol. Chem.
275
:
22114
22120
.
Bernatchez
,
G.
,
L.
Berrou
,
Z.
Benakezouh
,
J.
Ducay
, and
L.
Parent
.
2001
.
Role of Repeat I in the fast inactivation kinetics of the Ca(V)2.3 channel
.
Biochim. Biophys. Acta
.
1514
:
217
229
.
Berrou
,
L.
,
G.
Bernatchez
, and
L.
Parent
.
2001
.
Molecular determinants of inactivation within the I-II linker of alpha1E (CaV2.3) calcium channels
.
Biophys. J.
80
:
215
228
.
Bers
,
D.M.
1991
.
Ca regulation in cardiac muscle
.
Med. Sci. Sports Exerc.
23
:
1157
1162
.
Bers
,
D.M.
2002
.
Cardiac excitation-contraction coupling
.
Nature
.
415
:
198
205
.
Bezanilla
,
F.
, and
C.M.
Armstrong
.
1977
.
Inactivation of the sodium channel. I. Sodium current experiments
.
J. Gen. Physiol.
70
:
549
566
.
Bock
,
G.
,
M.
Gebhart
,
A.
Scharinger
,
W.
Jangsangthong
,
P.
Busquet
,
C.
Poggiani
,
S.
Sartori
,
M.E.
Mangoni
,
M.J.
Sinnegger-Brauns
,
S.
Herzig
, et al
.
2011
.
Functional properties of a newly identified C-terminal splice variant of Cav1.3 L-type Ca2+ channels
.
J. Biol. Chem.
286
:
42736
42748
.
Boczek
,
N.J.
,
E.M.
Miller
,
D.
Ye
,
V.V.
Nesterenko
,
D.J.
Tester
,
C.
Antzelevitch
,
R.J.
Czosek
,
M.J.
Ackerman
, and
S.M.
Ware
.
2015
.
Novel Timothy syndrome mutation leading to increase in CACNA1C window current
.
Heart Rhythm
.
12
:
211
219
.
Bourinet
,
E.
,
T.W.
Soong
,
K.
Sutton
,
S.
Slaymaker
,
E.
Mathews
,
A.
Monteil
,
G.W.
Zamponi
,
J.
Nargeot
, and
T.P.
Snutch
.
1999
.
Splicing of alpha 1A subunit gene generates phenotypic variants of P- and Q-type calcium channels
.
Nat. Neurosci.
2
:
407
415
.
Bozarth
,
X.
,
J.N.
Dines
,
Q.
Cong
,
G.M.
Mirzaa
,
K.
Foss
,
J.
Lawrence Merritt
II
,
J.
Thies
,
H.C.
Mefford
, and
E.
Novotny
.
2018
.
Expanding clinical phenotype in CACNA1C related disorders: From neonatal onset severe epileptic encephalopathy to late-onset epilepsy
.
Am. J. Med. Genet. A.
176
:
2733
2739
.
Brehm
,
P.
, and
R.
Eckert
.
1978
.
Calcium entry leads to inactivation of calcium channel in Paramecium
.
Science
.
202
:
1203
1206
.
Brum
,
G.
,
R.
Fitts
,
G.
Pizarro
, and
E.
Ríos
.
1988
.
Voltage sensors of the frog skeletal muscle membrane require calcium to function in excitation-contraction coupling
.
J. Physiol.
398
:
475
505
.
Bunda
,
A.
, and
A.
Andrade
.
2022
.
BaseScope™ approach to visualize alternative splice variants in tissue
.
Methods Mol. Biol.
2537
:
185
196
.
Buraei
,
Z.
, and
J.
Yang
.
2013
.
Structure and function of the β subunit of voltage-gated Ca²⁺ channels
.
Biochim. Biophys. Acta
.
1828
:
1530
1540
.
Cai
,
T.
,
H.I.
Jen
,
H.
Kang
,
T.J.
Klisch
,
H.Y.
Zoghbi
, and
A.K.
Groves
.
2015
.
Characterization of the transcriptome of nascent hair cells and identification of direct targets of the Atoh1 transcription factor
.
J. Neurosci.
35
:
5870
5883
.
Campiglio
,
M.
,
P.
Costé de Bagneaux
,
N.J.
Ortner
,
P.
Tuluc
,
F.
Van Petegem
, and
B.E.
Flucher
.
2018
.
STAC proteins associate to the IQ domain of CaV1.2 and inhibit calcium-dependent inactivation
.
Proc. Natl. Acad. Sci. USA
.
115
:
1376
1381
.
Catacuzzeno
,
L.
,
F.
Conti
, and
F.
Franciolini
.
2023
.
Fifty years of gating currents and channel gating
.
J. Gen. Physiol.
155
:e202313380.
Catterall
,
W.A.
2000
.
Structure and regulation of voltage-gated Ca2+ channels
.
Annu. Rev. Cell Dev. Biol.
16
:
521
555
.
Catterall
,
W.A.
,
E.
Perez-Reyes
,
T.P.
Snutch
, and
J.
Striessnig
.
2005
.
International Union of Pharmacology. XLVIII. Nomenclature and structure-function relationships of voltage-gated calcium channels
.
Pharmacol. Rev.
57
:
411
425
.
Cens
,
T.
,
S.
Restituito
, and
P.
Charnet
.
1999
.
Regulation of Ca-sensitive inactivation of a 1-type Ca2+ channel by specific domains of beta subunits
.
FEBS Lett.
450
:
17
22
.
Chaudhuri
,
D.
,
S.Y.
Chang
,
C.D.
DeMaria
,
R.S.
Alvania
,
T.W.
Soong
, and
D.T.
Yue
.
2004
.
Alternative splicing as a molecular switch for Ca2+/calmodulin-dependent facilitation of P/Q-type Ca2+ channels
.
J. Neurosci.
24
:
6334
6342
.
Chen
,
Z.
,
A.
Mondal
,
F.
Abderemane-Ali
,
S.
Jang
,
S.
Niranjan
,
J.L.
Montaño
,
B.W.
Zaro
, and
D.L.
Minor
Jr.
2023
.
EMC chaperone-CaV structure reveals an ion channel assembly intermediate
.
Nature
.
619
:
410
419
.
Choi
,
K.L.
,
R.W.
Aldrich
, and
G.
Yellen
.
1991
.
Tetraethylammonium blockade distinguishes two inactivation mechanisms in voltage-activated K+ channels
.
Proc. Natl. Acad. Sci. USA
.
88
:
5092
5095
.
Crotti
,
L.
,
C.N.
Johnson
,
E.
Graf
,
G.M.
De Ferrari
,
B.F.
Cuneo
,
M.
Ovadia
,
J.
Papagiannis
,
M.D.
Feldkamp
,
S.G.
Rathi
,
J.D.
Kunic
, et al
.
2013
.
Calmodulin mutations associated with recurrent cardiac arrest in infants
.
Circulation
.
127
:
1009
1017
.
Crotti
,
L.
,
C.
Spazzolini
,
D.J.
Tester
,
A.
Ghidoni
,
A.E.
Baruteau
,
B.M.
Beckmann
,
E.R.
Behr
,
J.S.
Bennett
,
C.R.
Bezzina
,
Z.A.
Bhuiyan
, et al
.
2019
.
Calmodulin mutations and life-threatening cardiac arrhythmias: Insights from the international calmodulinopathy registry
.
Eur. Heart J.
40
:
2964
2975
.
Cuenca-León
,
E.
,
I.
Banchs
,
S.A.
Serra
,
P.
Latorre
,
N.
Fernàndez-Castillo
,
R.
Corominas
,
M.A.
Valverde
,
V.
Volpini
,
J.M.
Fernández-Fernández
,
A.
Macaya
, and
B.
Cormand
.
2009
.
Late-onset episodic ataxia type 2 associated with a novel loss-of-function mutation in the CACNA1A gene
.
J. Neurol. Sci.
280
:
10
14
.
Cuttle
,
M.F.
,
T.
Tsujimoto
,
I.D.
Forsythe
, and
T.
Takahashi
.
1998
.
Facilitation of the presynaptic calcium current at an auditory synapse in rat brainstem
.
J. Physiol.
512
:
723
729
.
Dafi
,
O.
,
L.
Berrou
,
Y.
Dodier
,
A.
Raybaud
,
R.
Sauvé
, and
L.
Parent
.
2004
.
Negatively charged residues in the N-terminal of the AID helix confer slow voltage dependent inactivation gating to CaV1.2
.
Biophys. J.
87
:
3181
3192
.
Dannenberg
,
F.
,
A.
Von Moers
,
P.
Bittigau
,
J.
Lange
,
S.
Wiegand
,
F.
Török
,
G.
Stölting
,
J.
Striessnig
,
M.M.
Motazacker
,
M.F.
Broekema
, et al
.
2024
.
A novel de novo gain-of-function CACNA1D variant in neurodevelopmental disease with congenital tremor, seizures, and hypotonia
.
Neurol. Genet.
10
:e200186.
De Waard
,
M.
,
D.R.
Witcher
,
M.
Pragnell
,
H.
Liu
, and
K.P.
Campbell
.
1995
.
Properties of the alpha 1-beta anchoring site in voltage-dependent Ca2+ channels
.
J. Biol. Chem.
270
:
12056
12064
.
Del Rivero Morfin
,
P.J.
,
A.L.
Kochiss
,
K.R.
Liedl
,
B.E.
Flucher
,
M.L.I.
Fernández-Quintero
, and
M.
Ben-Johny
.
2024
.
Asymmetric contribution of a selectivity filter gate in triggering inactivation of CaV1.3 channels
.
J. Gen. Physiol.
156
:e202313365.
DeMaria
,
C.D.
,
T.W.
Soong
,
B.A.
Alseikhan
,
R.S.
Alvania
, and
D.T.
Yue
.
2001
.
Calmodulin bifurcates the local Ca2+ signal that modulates P/Q-type Ca2+ channels
.
Nature
.
411
:
484
489
.
Dick
,
I.E.
,
M.R.
Tadross
,
H.
Liang
,
L.H.
Tay
,
W.
Yang
, and
D.T.
Yue
.
2008
.
A modular switch for spatial Ca2+ selectivity in the calmodulin regulation of CaV channels
.
Nature
.
451
:
830
834
.
Dick
,
I.E.
,
R.
Joshi-Mukherjee
,
W.
Yang
, and
D.T.
Yue
.
2016
.
Arrhythmogenesis in Timothy Syndrome is associated with defects in Ca(2+)-dependent inactivation
.
Nat. Commun.
7
:
10370
.
Dong
,
Y.
,
Y.
Gao
,
S.
Xu
,
Y.
Wang
,
Z.
Yu
,
Y.
Li
,
B.
Li
,
T.
Yuan
,
B.
Yang
,
X.C.
Zhang
, et al
.
2021
.
Closed-state inactivation and pore-blocker modulation mechanisms of human CaV2.2
.
Cell Rep.
37
:
109931
.
Erickson
,
M.G.
,
B.A.
Alseikhan
,
B.Z.
Peterson
, and
D.T.
Yue
.
2001
.
Preassociation of calmodulin with voltage-gated Ca(2+) channels revealed by FRET in single living cells
.
Neuron
.
31
:
973
985
.
Erickson
,
M.G.
,
H.
Liang
,
M.X.
Mori
, and
D.T.
Yue
.
2003
.
FRET two-hybrid mapping reveals function and location of L-type Ca2+ channel CaM preassociation
.
Neuron
.
39
:
97
107
.
Ferreira
,
G.
,
E.
Ríos
, and
N.
Reyes
.
2003
.
Two components of voltage-dependent inactivation in Ca(v)1.2 channels revealed by its gating currents
.
Biophys. J.
84
:
3662
3678
.
Findeisen
,
F.
, and
D.L.
Minor
Jr.
2010
.
Structural basis for the differential effects of CaBP1 and calmodulin on Ca(V)1.2 calcium-dependent inactivation
.
Structure
.
18
:
1617
1631
.
Flucher
,
B.E.
2016
.
Specific contributions of the four voltage-sensing domains in L-type calcium channels to gating and modulation
.
J. Gen. Physiol.
148
:
91
95
.
Flucher
,
B.E.
, and
M.
Campiglio
.
2019
.
STAC proteins: The missing link in skeletal muscle EC coupling and new regulators of calcium channel function
.
Biochim. Biophys. Acta Mol. Cell Res.
1866
:
1101
1110
.
Gambeta
,
E.
,
M.A.
Gandini
,
I.A.
Souza
,
L.
Ferron
, and
G.W.
Zamponi
.
2021
.
A CACNA1A variant associated with trigeminal neuralgia alters the gating of Cav2.1 channels
.
Mol. Brain
.
14
:
4
.
Gao
,
S.
, and
N.
Yan
.
2021
.
Structural basis of the modulation of the voltage-gated calcium ion channel Cav 1.1 by dihydropyridine compounds*
.
Angew. Chem. Int. Ed. Engl.
60
:
3131
3137
.
Gao
,
S.
,
X.
Yao
,
J.
Chen
,
G.
Huang
,
X.
Fan
,
L.
Xue
,
Z.
Li
,
T.
Wu
,
Y.
Zheng
,
J.
Huang
, et al
.
2023
.
Structural basis for human Cav1.2 inhibition by multiple drugs and the neurotoxin calciseptine
.
Cell
.
186
:
5363
5374.e16
.
Gebhart
,
M.
,
G.
Juhasz-Vedres
,
A.
Zuccotti
,
N.
Brandt
,
J.
Engel
,
A.
Trockenbacher
,
G.
Kaur
,
G.J.
Obermair
,
M.
Knipper
,
A.
Koschak
, and
J.
Striessnig
.
2010
.
Modulation of Cav1.3 Ca2+ channel gating by Rab3 interacting molecule
.
Mol. Cell. Neurosci.
44
:
246
259
.
George
,
A.L.
Jr.
2015
.
Calmodulinopathy: A genetic trilogy
.
Heart Rhythm
.
12
:
423
424
.
Gillis
,
J.
,
E.
Burashnikov
,
C.
Antzelevitch
,
S.
Blaser
,
G.
Gross
,
L.
Turner
,
R.
Babul-Hirji
, and
D.
Chitayat
.
2012
.
Long QT, syndactyly, joint contractures, stroke and novel CACNA1C mutation: Expanding the spectrum of Timothy syndrome
.
Am. J. Med. Genet. A
.
158A
:
182
187
.
Gorman
,
K.M.
,
E.
Meyer
,
D.
Grozeva
,
E.
Spinelli
,
A.
McTague
,
A.
Sanchis-Juan
,
K.J.
Carss
,
E.
Bryant
,
A.
Reich
,
A.L.
Schneider
, et al
.
2019
.
Bi-allelic loss-of-function CACNA1B mutations in progressive epilepsy-dyskinesia
.
Am. J. Hum. Genet.
104
:
948
956
.
Griessmeier
,
K.
,
H.
Cuny
,
K.
Rötzer
,
O.
Griesbeck
,
H.
Harz
,
M.
Biel
, and
C.
Wahl-Schott
.
2009
.
Calmodulin is a functional regulator of Cav1.4 L-type Ca2+ channels
.
J. Biol. Chem.
284
:
29809
29816
.
Hardie
,
J.
, and
A.
Lee
.
2016
.
Decalmodulation of Cav1 channels by CaBPs
.
Channels
.
10
:
33
37
.
Helbig
,
K.L.
,
R.J.
Lauerer
,
J.C.
Bahr
,
I.A.
Souza
,
C.T.
Myers
,
B.
Uysal
,
N.
Schwarz
,
M.A.
Gandini
,
S.
Huang
,
B.
Keren
, et al
.
2018
.
De novo pathogenic variants in CACNA1E cause developmental and epileptic encephalopathy with contractures, macrocephaly, and dyskinesias
.
Am. J. Hum. Genet.
103
:
666
678
.
Helton
,
T.D.
,
W.
Xu
, and
D.
Lipscombe
.
2005
.
Neuronal L-type calcium channels open quickly and are inhibited slowly
.
J. Neurosci.
25
:
10247
10251
.
Hemara-Wahanui
,
A.
,
S.
Berjukow
,
C.I.
Hope
,
P.K.
Dearden
,
S.B.
Wu
,
J.
Wilson-Wheeler
,
D.M.
Sharp
,
P.
Lundon-Treweek
,
G.M.
Clover
,
J.C.
Hoda
, et al
.
2005
.
A CACNA1F mutation identified in an X-linked retinal disorder shifts the voltage dependence of Cav1.4 channel activation
.
Proc. Natl. Acad. Sci. USA
.
102
:
7553
7558
.
Hering
,
S.
,
S.
Aczél
,
R.L.
Kraus
,
S.
Berjukow
,
J.
Striessnig
, and
E.N.
Timin
.
1997
.
Molecular mechanism of use-dependent calcium channel block by phenylalkylamines: Role of inactivation
.
Proc. Natl. Acad. Sci. USA
.
94
:
13323
13328
.
Herlitze
,
S.
,
G.H.
Hockerman
,
T.
Scheuer
, and
W.A.
Catterall
.
1997
.
Molecular determinants of inactivation and G protein modulation in the intracellular loop connecting domains I and II of the calcium channel alpha1A subunit
.
Proc. Natl. Acad. Sci. USA
.
94
:
1512
1516
.
Herold
,
K.G.
,
J.W.
Hussey
, and
I.E.
Dick
.
2023
.
CACNA1C-related channelopathies
.
Handb. Exp. Pharmacol.
279
:
159
181
.
Hoda
,
J.C.
,
F.
Zaghetto
,
A.
Koschak
, and
J.
Striessnig
.
2005
.
Congenital stationary night blindness type 2 mutations S229P, G369D, L1068P, and W1440X alter channel gating or functional expression of Ca(v)1.4 L-type Ca2+ channels
.
J. Neurosci.
25
:
252
259
.
Hoda
,
J.C.
,
F.
Zaghetto
,
A.
Singh
,
A.
Koschak
, and
J.
Striessnig
.
2006
.
Effects of congenital stationary night blindness type 2 mutations R508Q and L1364H on Cav1.4 L-type Ca2+ channel function and expression
.
J. Neurochem.
96
:
1648
1658
.
Hodgkin
,
A.L.
,
A.F.
Huxley
, and
B.
Katz
.
1952
.
Measurement of current-voltage relations in the membrane of the giant axon of Loligo
.
J. Physiol.
116
:
424
448
.
Hofer
,
N.T.
,
P.
Tuluc
,
N.J.
Ortner
,
Y.V.
Nikonishyna
,
M.L.
Fernándes-Quintero
,
K.R.
Liedl
,
B.E.
Flucher
,
H.
Cox
, and
J.
Striessnig
.
2020
.
Biophysical classification of a CACNA1D de novo mutation as a high-risk mutation for a severe neurodevelopmental disorder
.
Mol. Autism
.
11
:
4
.
Hohaus
,
A.
,
S.
Beyl
,
M.
Kudrnac
,
S.
Berjukow
,
E.N.
Timin
,
R.
Marksteiner
,
M.A.
Maw
, and
S.
Hering
.
2005
.
Structural determinants of L-type channel activation in segment IIS6 revealed by a retinal disorder
.
J. Biol. Chem.
280
:
38471
38477
.
Hoppa
,
M.B.
,
B.
Lana
,
W.
Margas
,
A.C.
Dolphin
, and
T.A.
Ryan
.
2012
.
α2δ expression sets presynaptic calcium channel abundance and release probability
.
Nature
.
486
:
122
125
.
Hoshi
,
T.
,
W.N.
Zagotta
, and
R.W.
Aldrich
.
1990
.
Biophysical and molecular mechanisms of Shaker potassium channel inactivation
.
Science
.
250
:
533
538
.
Huang
,
H.
,
B.Z.
Tan
,
Y.
Shen
,
J.
Tao
,
F.
Jiang
,
Y.Y.
Sung
,
C.K.
Ng
,
M.
Raida
,
G.
Köhr
,
M.
Higuchi
, et al
.
2012
.
RNA editing of the IQ domain in Ca(v)1.3 channels modulates their Ca²⁺-dependent inactivation
.
Neuron
.
73
:
304
316
.
Huang
,
H.
,
K.
Kapeli
,
W.
Jin
,
Y.P.
Wong
,
T.V.
Arumugam
,
J.H.
Koh
,
S.
Srimasorn
,
K.
Mallilankaraman
,
J.J.E.
Chua
,
G.W.
Yeo
, and
T.W.
Soong
.
2018
.
Tissue-selective restriction of RNA editing of CaV1.3 by splicing factor SRSF9
.
Nucleic Acids Res.
46
:
7323
7338
.
Huang
,
J.
,
X.
Pan
, and
N.
Yan
.
2024
.
Structural biology and molecular pharmacology of voltage-gated ion channels
.
Nat. Rev. Mol. Cell Biol.
25
:
904
925
.
Hussey
,
J.W.
,
K.G.
Herold
, and
I.E.
Dick
.
2023a
.
Voltage-gated calcium channelopathies
. In
Ca Signals: From Single Molecules to Physiology
.
L.S.
Satin
,
M.
Ben-Johny
, and
I.E.
Dick
, editors. Vol. 4. First edition.
IOP ebooks
.
Bristol, UK
.
13-1
13-78
.
Hussey
,
J.W.
,
W.B.
Limpitikul
, and
I.E.
Dick
.
2023b
.
Calmodulin mutations in human disease
.
Channels
.
17
:
2165278
.
Hussey
,
J.W.
,
E.
DeMarco
,
D.
DiSilvestre
,
M.
Brohus
,
A.O.
Busuioc
,
E.D.
Iversen
,
H.H.
Jensen
,
M.
Nyegaard
,
M.T.
Overgaard
,
M.
Ben-Johny
, and
I.E.
Dick
.
2024
.
Voltage gated calcium channel dysregulation may contribute to neurological symptoms in calmodulinopathies
.
bioRxiv
.
(Preprint posted December 05, 2024)
.
Imredy
,
J.P.
, and
D.T.
Yue
.
1994
.
Mechanism of Ca(2+)-sensitive inactivation of L-type Ca2+ channels
.
Neuron
.
12
:
1301
1318
.
Inagaki
,
A.
, and
A.
Lee
.
2013
.
Developmental alterations in the biophysical properties of Ca(v) 1.3 Ca(2+) channels in mouse inner hair cells
.
Channels
.
7
:
171
181
.
Indriati
,
D.W.
,
N.
Kamasawa
,
K.
Matsui
,
A.L.
Meredith
,
M.
Watanabe
, and
R.
Shigemoto
.
2013
.
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.
33
:
3668
3678
.
Ivanina
,
T.
,
Y.
Blumenstein
,
E.
Shistik
,
R.
Barzilai
, and
N.
Dascal
.
2000
.
Modulation of L-type Ca2+ channels by gbeta gamma and calmodulin via interactions with N and C termini of alpha 1C
.
J. Biol. Chem.
275
:
39846
39854
.
Jacobs
,
A.
,
B.P.
Knight
,
K.T.
McDonald
, and
M.C.
Burke
.
2006
.
Verapamil decreases ventricular tachyarrhythmias in a patient with Timothy syndrome (LQT8)
.
Heart Rhythm
.
3
:
967
970
.
Jensen
,
H.H.
,
M.
Brohus
,
M.
Nyegaard
, and
M.T.
Overgaard
.
2018
.
Human calmodulin mutations
.
Front. Mol. Neurosci.
11
:
396
.
Johnson
,
B.D.
,
G.H.
Hockerman
,
T.
Scheuer
, and
W.A.
Catterall
.
1996
.
Distinct effects of mutations in transmembrane segment IVS6 on block of L-type calcium channels by structurally similar phenylalkylamines
.
Mol. Pharmacol.
50
:
1388
1400
.
Jones
,
L.P.
,
C.D.
DeMaria
, and
D.T.
Yue
.
1999
.
N-type calcium channel inactivation probed by gating-current analysis
.
Biophys. J.
76
:
2530
2552
.
Kanaya
,
S.
,
P.
Arlock
,
B.G.
Katzung
, and
L.M.
Hondeghem
.
1983
.
Diltiazem and verapamil preferentially block inactivated cardiac calcium channels
.
J. Mol. Cell. Cardiol.
15
:
145
148
.
Karagianni
,
K.
,
D.
Dafou
,
K.
Xanthopoulos
,
T.
Sklaviadis
, and
E.
Kanata
.
2024
.
RNA editing regulates glutamatergic synapses in the frontal cortex of a molecular subtype of Amyotrophic Lateral Sclerosis
.
Mol. Med.
30
:
101
.
Karagueuzian
,
H.S.
,
A.
Pezhouman
,
M.
Angelini
, and
R.
Olcese
.
2017
.
Enhanced late Na and Ca currents as effective antiarrhythmic drug targets
.
Front. Pharmacol.
8
:
36
.
Kink
,
J.A.
,
M.E.
Maley
,
R.R.
Preston
,
K.Y.
Ling
,
M.A.
Wallen-Friedman
,
Y.
Saimi
, and
C.
Kung
.
1990
.
Mutations in paramecium calmodulin indicate functional differences between the C-terminal and N-terminal lobes in vivo
.
Cell
.
62
:
165
174
.
Komuro
,
H.
, and
P.
Rakic
.
1992
.
Selective role of N-type calcium channels in neuronal migration
.
Science
.
257
:
806
809
.
Koschak
,
A.
,
D.
Reimer
,
D.
Walter
,
J.C.
Hoda
,
T.
Heinzle
,
M.
Grabner
, and
J.
Striessnig
.
2003
.
Cav1.4alpha1 subunits can form slowly inactivating dihydropyridine-sensitive L-type Ca2+ channels lacking Ca2+-dependent inactivation
.
J. Neurosci.
23
:
6041
6049
.
Kraus
,
R.L.
,
M.J.
Sinnegger
,
H.
Glossmann
,
S.
Hering
, and
J.
Striessnig
.
1998
.
Familial hemiplegic migraine mutations change alpha1A Ca2+ channel kinetics
.
J. Biol. Chem.
273
:
5586
5590
.
Landstrom
,
A.P.
,
N.J.
Boczek
,
D.
Ye
,
C.Y.
Miyake
,
C.M.
De la Uz
,
H.D.
Allen
,
M.J.
Ackerman
, and
J.J.
Kim
.
2016
.
Novel long QT syndrome-associated missense mutation, L762F, in CACNA1C-encoded L-type calcium channel imparts a slower inactivation tau and increased sustained and window current
.
Int. J. Cardiol.
220
:
290
298
.
Lee
,
K.S.
,
E.
Marban
, and
R.W.
Tsien
.
1985
.
Inactivation of calcium channels in mammalian heart cells: Joint dependence on membrane potential and intracellular calcium
.
J. Physiol.
364
:
395
411
.
Lee
,
A.
,
S.T.
Wong
,
D.
Gallagher
,
B.
Li
,
D.R.
Storm
,
T.
Scheuer
, and
W.A.
Catterall
.
1999
.
Ca2+/calmodulin binds to and modulates P/Q-type calcium channels
.
Nature
.
399
:
155
159
.
Lee
,
A.
,
T.
Scheuer
, and
W.A.
Catterall
.
2000
.
Ca2+/calmodulin-dependent facilitation and inactivation of P/Q-type Ca2+ channels
.
J. Neurosci.
20
:
6830
6838
.
Lee
,
A.
,
H.
Zhou
,
T.
Scheuer
, and
W.A.
Catterall
.
2003
.
Molecular determinants of Ca(2+)/calmodulin-dependent regulation of Ca(v)2.1 channels
.
Proc. Natl. Acad. Sci. USA
.
100
:
16059
16064
.
Legha
,
W.
,
S.
Gaillard
,
E.
Gascon
,
P.
Malapert
,
M.
Hocine
,
S.
Alonso
, and
A.
Moqrich
.
2010
.
stac1 and stac2 genes define discrete and distinct subsets of dorsal root ganglia neurons
.
Gene Expr. Patterns
.
10
:
368
375
.
Levy
,
R.J.
,
K.W.
Timothy
,
J.F.G.
Underwood
,
J.
Hall
,
J.A.
Bernstein
, and
S.P.
Paşca
.
2023
.
A cross-sectional study of the neuropsychiatric phenotype of CACNA1C-related disorder
.
Pediatr. Neurol.
138
:
101
106
.
Lewis
,
R.S.
, and
A.J.
Hudspeth
.
1983
.
Voltage- and ion-dependent conductances in solitary vertebrate hair cells
.
Nature
.
304
:
538
541
.
Liang
,
H.
,
C.D.
DeMaria
,
M.G.
Erickson
,
M.X.
Mori
,
B.A.
Alseikhan
, and
D.T.
Yue
.
2003
.
Unified mechanisms of Ca2+ regulation across the Ca2+ channel family
.
Neuron
.
39
:
951
960
.
Lieb
,
A.
,
A.
Scharinger
,
S.
Sartori
,
M.J.
Sinnegger-Brauns
, and
J.
Striessnig
.
2012
.
Structural determinants of CaV1.3 L-type calcium channel gating
.
Channels
.
6
:
197
205
.
Limpitikul
,
W.B.
,
I.E.
Dick
,
R.
Joshi-Mukherjee
,
M.T.
Overgaard
,
A.L.
George
Jr.
, and
D.T.
Yue
.
2014
.
Calmodulin mutations associated with long QT syndrome prevent inactivation of cardiac L-type Ca(2+) currents and promote proarrhythmic behavior in ventricular myocytes
.
J. Mol. Cell. Cardiol.
74
:
115
124
.
Limpitikul
,
W.B.
,
I.E.
Dick
,
M.
Ben-Johny
, and
D.T.
Yue
.
2016
.
An autism-associated mutation in CaV1.3 channels has opposing effects on voltage- and Ca(2+)-dependent regulation
.
Sci. Rep.
6
:
27235
.
Liu
,
X.
,
P.S.
Yang
,
W.
Yang
, and
D.T.
Yue
.
2010
.
Enzyme-inhibitor-like tuning of Ca(2+) channel connectivity with calmodulin
.
Nature
.
463
:
968
972
.
Livshitz
,
L.M.
, and
Y.
Rudy
.
2007
.
Regulation of Ca2+ and electrical alternans in cardiac myocytes: Role of CAMKII and repolarizing currents
.
Am. J. Physiol. Heart Circ. Physiol.
292
:
H2854
H2866
.
Lu
,
Z.J.
,
A.
Pereverzev
,
H.L.
Liu
,
M.
Weiergräber
,
M.
Henry
,
A.
Krieger
,
N.
Smyth
,
J.
Hescheler
, and
T.
Schneider
.
2004
.
Arrhythmia in isolated prenatal hearts after ablation of the Cav2.3 (alpha1E) subunit of voltage-gated Ca2+ channels
.
Cell. Physiol. Biochem.
14
:
11
22
.
Ma
,
H.
,
R.D.
Groth
,
D.G.
Wheeler
,
C.F.
Barrett
, and
R.W.
Tsien
.
2011
.
Excitation-transcription coupling in sympathetic neurons and the molecular mechanism of its initiation
.
Neurosci. Res.
70
:
2
8
.
Marcantoni
,
A.
,
C.
Calorio
,
E.
Hidisoglu
,
G.
Chiantia
, and
E.
Carbone
.
2020
.
Cav1.2 channelopathies causing autism: New hallmarks on Timothy syndrome
.
Pflugers Arch.
472
:
775
789
.
McDavid
,
S.
, and
K.P.
Currie
.
2006
.
G-proteins modulate cumulative inactivation of N-type (Cav2.2) calcium channels
.
J. Neurosci.
26
:
13373
13383
.
McRory
,
J.E.
,
J.
Hamid
,
C.J.
Doering
,
E.
Garcia
,
R.
Parker
,
K.
Hamming
,
L.
Chen
,
M.
Hildebrand
,
A.M.
Beedle
,
L.
Feldcamp
, et al
.
2004
.
The CACNA1F gene encodes an L-type calcium channel with unique biophysical properties and tissue distribution
.
J. Neurosci.
24
:
1707
1718
.
Mochida
,
S.
,
A.P.
Few
,
T.
Scheuer
, and
W.A.
Catterall
.
2008
.
Regulation of presynaptic Ca(V)2.1 channels by Ca2+ sensor proteins mediates short-term synaptic plasticity
.
Neuron
.
57
:
210
216
.
Morales
,
D.
,
T.
Hermosilla
, and
D.
Varela
.
2019
.
Calcium-dependent inactivation controls cardiac L-type Ca2+ currents under β-adrenergic stimulation
.
J. Gen. Physiol.
151
:
786
797
.
Mori
,
M.X.
,
C.W.
Vander Kooi
,
D.J.
Leahy
, and
D.T.
Yue
.
2008
.
Crystal structure of the CaV2 IQ domain in complex with Ca2+/calmodulin: High-resolution mechanistic implications for channel regulation by Ca2+
.
Structure
.
16
:
607
620
.
Minor
,
D.L.
Jr.
, and
F.
Findeisen
.
2010
.
Progress in the structural understanding of voltage-gated calcium channel (CaV) function and modulation
.
Channels
.
4
:
459
474
.
Morotti
,
S.
,
E.
Grandi
,
A.
Summa
,
K.S.
Ginsburg
, and
D.M.
Bers
.
2012
.
Theoretical study of L-type Ca(2+) current inactivation kinetics during action potential repolarization and early afterdepolarizations
.
J. Physiol.
590
:
4465
4481
.
Müller
,
C.S.
,
A.
Haupt
,
W.
Bildl
,
J.
Schindler
,
H.G.
Knaus
,
M.
Meissner
,
B.
Rammner
,
J.
Striessnig
,
V.
Flockerzi
,
B.
Fakler
, and
U.
Schulte
.
2010
.
Quantitative proteomics of the Cav2 channel nano-environments in the mammalian brain
.
Proc. Natl. Acad. Sci. USA
.
107
:
14950
14957
.
Nanou
,
E.
, and
W.A.
Catterall
.
2018
.
Calcium channels, synaptic plasticity, and neuropsychiatric disease
.
Neuron
.
98
:
466
481
.
Nanou
,
E.
,
T.
Scheuer
, and
W.A.
Catterall
.
2016
.
Calcium sensor regulation of the CaV2.1 Ca2+ channel contributes to long-term potentiation and spatial learning
.
Proc. Natl. Acad. Sci. USA
.
113
:
13209
13214
.
Nelson
,
B.R.
,
F.
Wu
,
Y.
Liu
,
D.M.
Anderson
,
J.
McAnally
,
W.
Lin
,
S.C.
Cannon
,
R.
Bassel-Duby
, and
E.N.
Olson
.
2013
.
Skeletal muscle-specific T-tubule protein STAC3 mediates voltage-induced Ca2+ release and contractility
.
Proc. Natl. Acad. Sci. USA
.
110
:
11881
11886
.
Nilius
,
B.
,
P.
Hess
,
J.B.
Lansman
, and
R.W.
Tsien
.
1986
.
A novel type of cardiac calcium channel in ventricular cells
.
Biomed. Biochim. Acta
.
45
:
S167
S170
.
Niu
,
J.
,
I.E.
Dick
,
W.
Yang
,
M.A.
Bamgboye
,
D.T.
Yue
,
G.
Tomaselli
,
T.
Inoue
, and
M.
Ben-Johny
.
2018a
.
Allosteric regulators selectively prevent Ca2+-feedback of CaV and NaV channels
.
Elife
.
7
:e35222.
Niu
,
J.
,
W.
Yang
,
D.T.
Yue
,
T.
Inoue
, and
M.
Ben-Johny
.
2018b
.
Duplex signaling by CaM and Stac3 enhances CaV1.1 function and provides insights into congenital myopathy
.
J. Gen. Physiol.
150
:
1145
1161
.
Nowycky
,
M.C.
,
A.P.
Fox
, and
R.W.
Tsien
.
1985
.
Three types of neuronal calcium channel with different calcium agonist sensitivity
.
Nature
.
316
:
440
443
.
Nyegaard
,
M.
, and
M.T.
Overgaard
.
2019
.
The international calmodulinopathy registry: Recording the diverse phenotypic spectrum of un-CALM hearts
.
Eur. Heart J.
40
:
2976
2978
.
Olcese
,
R.
,
N.
Qin
,
T.
Schneider
,
A.
Neely
,
X.
Wei
,
E.
Stefani
, and
L.
Birnbaumer
.
1994
.
The amino terminus of a calcium channel beta subunit sets rates of channel inactivation independently of the subunit’s effect on activation
.
Neuron
.
13
:
1433
1438
.
Ortner
,
N.J.
,
T.
Kaserer
,
J.N.
Copeland
, and
J.
Striessnig
.
2020a
.
De novo CACNA1D Ca2+ channelopathies: Clinical phenotypes and molecular mechanism
.
Pflugers Arch.
472
:
755
773
.
Ortner
,
N.J.
,
A.
Pinggera
,
N.T.
Hofer
,
A.
Siller
,
N.
Brandt
,
A.
Raffeiner
,
K.
Vilusic
,
I.
Lang
,
K.
Blum
,
G.J.
Obermair
, et al
.
2020b
.
RBP2 stabilizes slow Cav1.3 Ca2+ channel inactivation properties of cochlear inner hair cells
.
Pflugers Arch.
472
:
3
25
.
Ortner
,
N.J.
,
A.
Sah
,
E.
Paradiso
,
J.
Shin
,
S.
Stojanovic
,
N.
Hammer
,
M.
Haritonova
,
N.T.
Hofer
,
A.
Marcantoni
,
L.
Guarina
, et al
.
2023
.
The human channel gating-modifying A749G CACNA1D (Cav1.3) variant induces a neurodevelopmental syndrome-like phenotype in mice
.
JCI Insight
.
8
:e162100.
Pajouhesh
,
H.
,
Z.P.
Feng
,
Y.
Ding
,
L.
Zhang
,
H.
Pajouhesh
,
J.L.
Morrison
,
F.
Belardetti
,
E.
Tringham
,
E.
Simonson
,
T.W.
Vanderah
, et al
.
2010
.
Structure-activity relationships of diphenylpiperazine N-type calcium channel inhibitors
.
Bioorg. Med. Chem. Lett.
20
:
1378
1383
.
Patil
,
P.G.
,
D.L.
Brody
, and
D.T.
Yue
.
1998
.
Preferential closed-state inactivation of neuronal calcium channels
.
Neuron
.
20
:
1027
1038
.
Peterson
,
B.Z.
,
C.D.
DeMaria
,
J.P.
Adelman
, and
D.T.
Yue
.
1999
.
Calmodulin is the Ca2+ sensor for Ca2+ -dependent inactivation of L-type calcium channels
.
Neuron
.
22
:
549
558
.
Pinggera
,
A.
, and
J.
Striessnig
.
2016
.
Cav 1.3 (CACNA1D) L-type Ca2+ channel dysfunction in CNS disorders
.
J. Physiol.
594
:
5839
5849
.
Pinggera
,
A.
,
A.
Lieb
,
B.
Benedetti
,
M.
Lampert
,
S.
Monteleone
,
K.R.
Liedl
,
P.
Tuluc
, and
J.
Striessnig
.
2015
.
CACNA1D de novo mutations in autism spectrum disorders activate Cav1.3 L-type calcium channels
.
Biol. Psychiatry
.
77
:
816
822
.
Pinggera
,
A.
,
L.
Mackenroth
,
A.
Rump
,
J.
Schallner
,
F.
Beleggia
,
B.
Wollnik
, and
J.
Striessnig
.
2017
.
New gain-of-function mutation shows CACNA1D as recurrently mutated gene in autism spectrum disorders and epilepsy
.
Hum. Mol. Genet.
26
:
2923
2932
.
Pitt
,
G.S.
,
R.D.
Zühlke
,
A.
Hudmon
,
H.
Schulman
,
H.
Reuter
, and
R.W.
Tsien
.
2001
.
Molecular basis of calmodulin tethering and Ca2+-dependent inactivation of L-type Ca2+ channels
.
J. Biol. Chem.
276
:
30794
30802
.
Pitt
,
G.S.
,
M.
Matsui
, and
C.
Cao
.
2021
.
Voltage-gated calcium channels in nonexcitable tissues
.
Annu. Rev. Physiol.
83
:
183
203
.
Pizarro
,
G.
,
R.
Fitts
,
I.
Uribe
, and
E.
Ríos
.
1989
.
The voltage sensor of excitation-contraction coupling in skeletal muscle. Ion dependence and selectivity
.
J. Gen. Physiol.
94
:
405
428
.
Polster
,
A.
,
S.
Perni
,
H.
Bichraoui
, and
K.G.
Beam
.
2015
.
Stac adaptor proteins regulate trafficking and function of muscle and neuronal L-type Ca2+ channels
.
Proc. Natl. Acad. Sci. USA
.
112
:
602
606
.
Polster
,
A.
,
P.J.
Dittmer
,
S.
Perni
,
H.
Bichraoui
,
W.A.
Sather
, and
K.G.
Beam
.
2018
.
Stac proteins suppress Ca2+-dependent inactivation of neuronal l-type Ca2+ channels
.
J. Neurosci.
38
:
9215
9227
.
Raybaud
,
A.
,
Y.
Dodier
,
P.
Bissonnette
,
M.
Simoes
,
D.G.
Bichet
,
R.
Sauvé
, and
L.
Parent
.
2006
.
The role of the GX9GX3G motif in the gating of high voltage-activated Ca2+ channels
.
J. Biol. Chem.
281
:
39424
39436
.
Retterer
,
K.
,
J.
Juusola
,
M.T.
Cho
,
P.
Vitazka
,
F.
Millan
,
F.
Gibellini
,
A.
Vertino-Bell
,
N.
Smaoui
,
J.
Neidich
,
K.G.
Monaghan
, et al
.
2016
.
Clinical application of whole-exome sequencing across clinical indications
.
Genet. Med.
18
:
696
704
.
Reuter
,
H.
,
C.F.
Stevens
,
R.W.
Tsien
, and
G.
Yellen
.
1982
.
Properties of single calcium channels in cardiac cell culture
.
Nature
.
297
:
501
504
.
Rodan
,
L.H.
,
R.C.
Spillmann
,
H.T.
Kurata
,
S.M.
Lamothe
,
J.
Maghera
,
R.A.
Jamra
,
A.
Alkelai
,
S.E.
Antonarakis
,
I.
Atallah
,
O.
Bar-Yosef
, et al
.
2021
.
Phenotypic expansion of CACNA1C-associated disorders to include isolated neurological manifestations
.
Genet. Med.
23
:
1922
1932
.
Sampedro-Castañeda
,
M.
,
L.L.
Baltussen
,
A.T.
Lopes
,
Y.
Qiu
,
L.
Sirvio
,
S.R.
Mihaylov
,
S.
Claxton
,
J.C.
Richardson
,
G.
Lignani
, and
S.K.
Ultanir
.
2023
.
Epilepsy-linked kinase CDKL5 phosphorylates voltage-gated calcium channel Cav2.3, altering inactivation kinetics and neuronal excitability
.
Nat. Commun.
14
:
7830
.
Sang
,
L.
,
I.E.
Dick
, and
D.T.
Yue
.
2016
.
Protein kinase A modulation of CaV1.4 calcium channels
.
Nat. Commun.
7
:
12239
.
Sang
,
L.
,
D.C.O.
Vieira
,
D.T.
Yue
,
M.
Ben-Johny
, and
I.E.
Dick
.
2021
.
The molecular basis of the inhibition of CaV1 calcium-dependent inactivation by the distal carboxy tail
.
J. Biol. Chem.
296
:
100502
.
Sather
,
W.A.
,
J.
Yang
, and
R.W.
Tsien
.
1994
.
Structural basis of ion channel permeation and selectivity
.
Curr. Opin. Neurobiol.
4
:
313
323
.
Scharinger
,
A.
,
S.
Eckrich
,
D.H.
Vandael
,
K.
Schönig
,
A.
Koschak
,
D.
Hecker
,
G.
Kaur
,
A.
Lee
,
A.
Sah
,
D.
Bartsch
, et al
.
2015
.
Cell-type-specific tuning of Cav1.3 Ca(2+)-channels by a C-terminal automodulatory domain
.
Front. Cell. Neurosci.
9
:
309
.
Schneider
,
M.F.
, and
W.K.
Chandler
.
1973
.
Voltage dependent charge movement of skeletal muscle: A possible step in excitation-contraction coupling
.
Nature
.
242
:
244
246
.
Serra
,
S.A.
,
N.
Fernàndez-Castillo
,
A.
Macaya
,
B.
Cormand
,
M.A.
Valverde
, and
J.M.
Fernández-Fernández
.
2009
.
The hemiplegic migraine-associated Y1245C mutation in CACNA1A results in a gain of channel function due to its effect on the voltage sensor and G-protein-mediated inhibition
.
Pflugers Arch.
458
:
489
502
.
Shah
,
D.P.
,
J.L.
Baez-Escudero
,
I.L.
Weisberg
,
J.F.
Beshai
, and
M.C.
Burke
.
2012
.
Ranolazine safely decreases ventricular and atrial fibrillation in Timothy syndrome (LQT8)
.
Pacing Clin. Electrophysiol.
35
:
e62
e64
.
Sheng
,
X.
,
T.
Nakada
,
M.
Kobayashi
,
T.
Kashihara
,
T.
Shibazaki
,
M.
Horiuchi-Hirose
,
S.
Gomi
,
M.
Hirose
,
T.
Aoyama
, and
M.
Yamada
.
2012
.
Two mechanistically distinct effects of dihydropyridine nifedipine on CaV1.2 L-type Ca²⁺ channels revealed by Timothy syndrome mutation
.
Eur. J. Pharmacol.
685
:
15
23
.
Simms
,
B.A.
, and
G.W.
Zamponi
.
2014
.
Neuronal voltage-gated calcium channels: Structure, function, and dysfunction
.
Neuron
.
82
:
24
45
.
Singh
,
A.
,
D.
Hamedinger
,
J.C.
Hoda
,
M.
Gebhart
,
A.
Koschak
,
C.
Romanin
, and
J.
Striessnig
.
2006
.
C-terminal modulator controls Ca2+-dependent gating of Ca(v)1.4 L-type Ca2+ channels
.
Nat. Neurosci.
9
:
1108
1116
.
Singh
,
A.
,
M.
Gebhart
,
R.
Fritsch
,
M.J.
Sinnegger-Brauns
,
C.
Poggiani
,
J.C.
Hoda
,
J.
Engel
,
C.
Romanin
,
J.
Striessnig
, and
A.
Koschak
.
2008
.
Modulation of voltage- and Ca2+-dependent gating of CaV1.3 L-type calcium channels by alternative splicing of a C-terminal regulatory domain
.
J. Biol. Chem.
283
:
20733
20744
.
Song
,
L.
,
S.E.
Park
,
Y.
Isseroff
,
K.
Morikawa
, and
M.
Yazawa
.
2017
.
Inhibition of CDK5 alleviates the cardiac phenotypes in Timothy syndrome
.
Stem Cell Rep.
9
:
50
57
.
Soong
,
T.W.
,
C.D.
DeMaria
,
R.S.
Alvania
,
L.S.
Zweifel
,
M.C.
Liang
,
S.
Mittman
,
W.S.
Agnew
, and
D.T.
Yue
.
2002
.
Systematic identification of splice variants in human P/Q-type channel alpha1(2.1) subunits: Implications for current density and Ca2+-dependent inactivation
.
J. Neurosci.
22
:
10142
10152
.
Spacey
,
S.D.
,
M.E.
Hildebrand
,
L.A.
Materek
,
T.D.
Bird
, and
T.P.
Snutch
.
2004
.
Functional implications of a novel EA2 mutation in the P/Q-type calcium channel
.
Ann. Neurol.
56
:
213
220
.
Splawski
,
I.
,
K.W.
Timothy
,
L.M.
Sharpe
,
N.
Decher
,
P.
Kumar
,
R.
Bloise
,
C.
Napolitano
,
P.J.
Schwartz
,
R.M.
Joseph
,
K.
Condouris
, et al
.
2004
.
Ca(V)1.2 calcium channel dysfunction causes a multisystem disorder including arrhythmia and autism
.
Cell
.
119
:
19
31
.
Splawski
,
I.
,
K.W.
Timothy
,
N.
Decher
,
P.
Kumar
,
F.B.
Sachse
,
A.H.
Beggs
,
M.C.
Sanguinetti
, and
M.T.
Keating
.
2005
.
Severe arrhythmia disorder caused by cardiac L-type calcium channel mutations
.
Proc. Natl. Acad. Sci. USA
.
102
:
8089
8096; discussion 8086–8088
.
Stephens
,
G.J.
,
K.M.
Page
,
Y.
Bogdanov
, and
A.C.
Dolphin
.
2000
.
The alpha1B Ca2+ channel amino terminus contributes determinants for beta subunit-mediated voltage-dependent inactivation properties
.
J. Physiol.
525
:
377
390
.
Stephens
,
R.F.
,
W.
Guan
,
B.S.
Zhorov
, and
J.D.
Spafford
.
2015
.
Selectivity filters and cysteine-rich extracellular loops in voltage-gated sodium, calcium, and NALCN channels
.
Front. Physiol.
6
:
153
.
Stern
,
M.D.
1992
.
Buffering of calcium in the vicinity of a channel pore
.
Cell Calcium
.
13
:
183
192
.
Stotz
,
S.C.
, and
G.W.
Zamponi
.
2001a
.
Identification of inactivation determinants in the domain IIS6 region of high voltage-activated calcium channels
.
J. Biol. Chem.
276
:
33001
33010
.
Stotz
,
S.C.
, and
G.W.
Zamponi
.
2001b
.
Structural determinants of fast inactivation of high voltage-activated Ca(2+) channels
.
Trends Neurosci.
24
:
176
181
.
Stotz
,
S.C.
,
J.
Hamid
,
R.L.
Spaetgens
,
S.E.
Jarvis
, and
G.W.
Zamponi
.
2000
.
Fast inactivation of voltage-dependent calcium channels. A hinged-lid mechanism?
J. Biol. Chem.
275
:
24575
24582
.
Stotz
,
S.C.
,
S.E.
Jarvis
, and
G.W.
Zamponi
.
2004
.
Functional roles of cytoplasmic loops and pore lining transmembrane helices in the voltage-dependent inactivation of HVA calcium channels
.
J. Physiol.
554
:
263
273
.
Striessnig
,
J.
,
A.
Koschak
,
M.J.
Sinnegger-Brauns
,
A.
Hetzenauer
,
N.K.
Nguyen
,
P.
Busquet
,
G.
Pelster
, and
N.
Singewald
.
2006
.
Role of voltage-gated L-type Ca2+ channel isoforms for brain function
.
Biochem. Soc. Trans.
34
:
903
909
.
Suzuki
,
H.
,
J.
Kawai
,
C.
Taga
,
T.
Yaoi
,
A.
Hara
,
K.
Hirose
,
Y.
Hayashizaki
, and
S.
Watanabe
.
1996
.
Stac, a novel neuron-specific protein with cysteine-rich and SH3 domains
.
Biochem. Biophys. Res. Commun.
229
:
902
909
.
Swensen
,
A.M.
,
J.
Herrington
,
R.M.
Bugianesi
,
G.
Dai
,
R.J.
Haedo
,
K.S.
Ratliff
,
M.M.
Smith
,
V.A.
Warren
,
S.P.
Arneric
,
C.
Eduljee
, et al
.
2012
.
Characterization of the substituted N-triazole oxindole TROX-1, a small-molecule, state-dependent inhibitor of Ca(V)2 calcium channels
.
Mol. Pharmacol.
81
:
488
497
.
Tadross
,
M.R.
,
I.E.
Dick
, and
D.T.
Yue
.
2008
.
Mechanism of local and global Ca2+ sensing by calmodulin in complex with a Ca2+ channel
.
Cell
.
133
:
1228
1240
.
Tadross
,
M.R.
,
M.
Ben Johny
, and
D.T.
Yue
.
2010
.
Molecular endpoints of Ca2+/calmodulin- and voltage-dependent inactivation of Ca(v)1.3 channels
.
J. Gen. Physiol.
135
:
197
215
.
Takahashi
,
K.
,
T.
Ishikawa
,
N.
Makita
,
K.
Takefuta
,
T.
Nabeshima
, and
M.
Nakayashiro
.
2016
.
A novel de novo calmodulin mutation in a 6-year-old boy who experienced an aborted cardiac arrest
.
HeartRhythm Case Rep.
3
:
69
72
.
Tan
,
B.Z.
,
F.
Jiang
,
M.Y.
Tan
,
D.
Yu
,
H.
Huang
,
Y.
Shen
, and
T.W.
Soong
.
2011
.
Functional characterization of alternative splicing in the C terminus of L-type CaV1.3 channels
.
J. Biol. Chem.
286
:
42725
42735
.
Tan
,
G.M.
,
D.
Yu
,
J.
Wang
, and
T.W.
Soong
.
2012
.
Alternative splicing at C terminus of Ca(V)1.4 calcium channel modulates calcium-dependent inactivation, activation potential, and current density
.
J. Biol. Chem.
287
:
832
847
.
Timothy
,
K.W.
,
R.
Bauer
,
K.A.
Larkin
,
E.P.
Walsh
,
D.J.
Abrams
,
C.
Gonzalez Corcia
,
A.
Valsamakis
,
G.S.
Pitt
,
I.E.
Dick
, and
A.
Golden
.
2024
.
A natural history study of Timothy Syndrome
.
Orphanet J. Rare Dis.
19
:
433
.
Tottene
,
A.
,
F.
Pivotto
,
T.
Fellin
,
T.
Cesetti
,
A.M.
van den Maagdenberg
, and
D.
Pietrobon
.
2005
.
Specific kinetic alterations of human CaV2.1 calcium channels produced by mutation S218L causing familial hemiplegic migraine and delayed cerebral edema and coma after minor head trauma
.
J. Biol. Chem.
280
:
17678
17686
.
Tsujimoto
,
T.
,
A.
Jeromin
,
N.
Saitoh
,
J.C.
Roder
, and
T.
Takahashi
.
2002
.
Neuronal calcium sensor 1 and activity-dependent facilitation of P/Q-type calcium currents at presynaptic nerve terminals
.
Science
.
295
:
2276
2279
.
Tuinte
,
W.E.
,
E.
Török
,
I.
Mahlknecht
,
P.
Tuluc
,
B.E.
Flucher
, and
M.
Campiglio
.
2022
.
STAC3 determines the slow activation kinetics of CaV 1.1 currents and inhibits its voltage-dependent inactivation
.
J. Cell. Physiol.
237
:
4197
4214
.
Van Petegem
,
F.
,
F.C.
Chatelain
, and
D.L.
Minor
Jr.
2005
.
Insights into voltage-gated calcium channel regulation from the structure of the CaV1.2 IQ domain-Ca2+/calmodulin complex
.
Nat. Struct. Mol. Biol.
12
:
1108
1115
.
Villalba-Galea
,
C.A.
,
W.
Sandtner
,
D.M.
Starace
, and
F.
Bezanilla
.
2008
.
S4-based voltage sensors have three major conformations
.
Proc. Natl. Acad. Sci. USA
.
105
:
17600
17607
.
Wahl-Schott
,
C.
,
L.
Baumann
,
H.
Cuny
,
C.
Eckert
,
K.
Griessmeier
, and
M.
Biel
.
2006
.
Switching off calcium-dependent inactivation in L-type calcium channels by an autoinhibitory domain
.
Proc. Natl. Acad. Sci. USA
.
103
:
15657
15662
.
Webster
,
G.
,
Z.J.
Schoppen
, and
A.L.
George
Jr.
2017
.
Treatment of calmodulinopathy with verapamil
.
BMJ Case Rep.
2017
:
bcr2017220568
.
Weiss
,
N.
,
A.
Sandoval
,
S.
Kyonaka
,
R.
Felix
,
Y.
Mori
, and
M.
De Waard
.
2011
.
Rim1 modulates direct G-protein regulation of Ca(v)2.2 channels
.
Pflugers Arch.
461
:
447
459
.
Wemhöner
,
K.
,
C.
Friedrich
,
B.
Stallmeyer
,
A.J.
Coffey
,
A.
Grace
,
S.
Zumhagen
,
G.
Seebohm
,
B.
Ortiz-Bonnin
,
S.
Rinné
,
F.B.
Sachse
, et al
.
2015
.
Gain-of-function mutations in the calcium channel CACNA1C (Cav1.2) cause non-syndromic long-QT but not Timothy syndrome
.
J. Mol. Cell. Cardiol.
80
:
186
195
.
West
,
J.W.
,
D.E.
Patton
,
T.
Scheuer
,
Y.
Wang
,
A.L.
Goldin
, and
W.A.
Catterall
.
1992
.
A cluster of hydrophobic amino acid residues required for fast Na(+)-channel inactivation
.
Proc. Natl. Acad. Sci. USA
.
89
:
10910
10914
.
Wilhelm
,
M.
,
J.
Schlegl
,
H.
Hahne
,
A.M.
Gholami
,
M.
Lieberenz
,
M.M.
Savitski
,
E.
Ziegler
,
L.
Butzmann
,
S.
Gessulat
,
H.
Marx
, et al
.
2014
.
Mass-spectrometry-based draft of the human proteome
.
Nature
.
509
:
582
587
.
Williams
,
B.
,
F.
Haeseleer
, and
A.
Lee
.
2018
.
Splicing of an automodulatory domain in Cav1.4 Ca2+ channels confers distinct regulation by calmodulin
.
J. Gen. Physiol.
150
:
1676
1687
.
Wong King Yuen
,
S.M.
,
M.
Campiglio
,
C.C.
Tung
,
B.E.
Flucher
, and
F.
Van Petegem
.
2017
.
Structural insights into binding of STAC proteins to voltage-gated calcium channels
.
Proc. Natl. Acad. Sci. USA
.
114
:
E9520
E9528
.
Wormuth
,
C.
,
A.
Lundt
,
C.
Henseler
,
R.
Müller
,
K.
Broich
,
A.
Papazoglou
, and
M.
Weiergräber
.
2016
.
Review: Cav2.3 R-type voltage-gated Ca2+ channels - functional implications in convulsive and non-convulsive seizure activity
.
Open Neurol. J.
10
:
99
126
.
Wu
,
L.G.
,
R.E.
Westenbroek
,
J.G.
Borst
,
W.A.
Catterall
, and
B.
Sakmann
.
1999
.
Calcium channel types with distinct presynaptic localization couple differentially to transmitter release in single calyx-type synapses
.
J. Neurosci.
19
:
726
736
.
Wu
,
J.
,
Z.
Yan
,
Z.
Li
,
C.
Yan
,
S.
Lu
,
M.
Dong
, and
N.
Yan
.
2015
.
Structure of the voltage-gated calcium channel Cav1.1 complex
.
Science
.
350
:
aad2395
.
Xia
,
X.M.
,
B.
Fakler
,
A.
Rivard
,
G.
Wayman
,
T.
Johnson-Pais
,
J.E.
Keen
,
T.
Ishii
,
B.
Hirschberg
,
C.T.
Bond
,
S.
Lutsenko
, et al
.
1998
.
Mechanism of calcium gating in small-conductance calcium-activated potassium channels
.
Nature
.
395
:
503
507
.
Xie
,
C.
,
X.G.
Zhen
, and
J.
Yang
.
2005
.
Localization of the activation gate of a voltage-gated Ca2+ channel
.
J. Gen. Physiol.
126
:
205
212
.
Xu
,
W.
, and
D.
Lipscombe
.
2001
.
Neuronal Ca(V)1.3alpha(1) L-type channels activate at relatively hyperpolarized membrane potentials and are incompletely inhibited by dihydropyridines
.
J. Neurosci.
21
:
5944
5951
.
Xu
,
J.
, and
L.G.
Wu
.
2005
.
The decrease in the presynaptic calcium current is a major cause of short-term depression at a calyx-type synapse
.
Neuron
.
46
:
633
645
.
Yang
,
P.S.
,
B.A.
Alseikhan
,
H.
Hiel
,
L.
Grant
,
M.X.
Mori
,
W.
Yang
,
P.A.
Fuchs
, and
D.T.
Yue
.
2006
.
Switching of Ca2+-dependent inactivation of Ca(v)1.3 channels by calcium binding proteins of auditory hair cells
.
J. Neurosci.
26
:
10677
10689
.
Yang
,
P.S.
,
M.B.
Johny
, and
D.T.
Yue
.
2014
.
Allostery in Ca²⁺ channel modulation by calcium-binding proteins
.
Nat. Chem. Biol.
10
:
231
238
.
Yao
,
X.
,
S.
Gao
,
J.
Wang
,
Z.
Li
,
J.
Huang
,
Y.
Wang
,
Z.
Wang
,
J.
Chen
,
X.
Fan
,
W.
Wang
, et al
.
2022
.
Structural basis for the severe adverse interaction of sofosbuvir and amiodarone on L-type Cav channels
.
Cell
.
185
:
4801
4810.e13
.
Yarotskyy
,
V.
, and
K.S.
Elmslie
.
2007
.
Roscovitine, a cyclin-dependent kinase inhibitor, affects several gating mechanisms to inhibit cardiac L-type (Ca(V)1.2) calcium channels
.
Br. J. Pharmacol.
152
:
386
395
.
Yarotskyy
,
V.
,
G.
Gao
,
B.Z.
Peterson
, and
K.S.
Elmslie
.
2009
.
The Timothy syndrome mutation of cardiac CaV1.2 (L-type) channels: Multiple altered gating mechanisms and pharmacological restoration of inactivation
.
J. Physiol.
587
:
551
565
.
Yarotskyy
,
V.
,
G.
Gao
,
L.
Du
,
S.B.
Ganapathi
,
B.Z.
Peterson
, and
K.S.
Elmslie
.
2010
.
Roscovitine binds to novel L-channel (CaV1.2) sites that separately affect activation and inactivation
.
J. Biol. Chem.
285
:
43
53
.
Yue
,
D.T.
,
P.H.
Backx
, and
J.P.
Imredy
.
1990
.
Calcium-sensitive inactivation in the gating of single calcium channels
.
Science
.
250
:
1735
1738
.
Zamponi
,
G.W.
,
J.
Striessnig
,
A.
Koschak
, and
A.C.
Dolphin
.
2015
.
The physiology, pathology, and pharmacology of voltage-gated calcium channels and their future therapeutic potential
.
Pharmacol. Rev.
67
:
821
870
.
Zhai
,
J.
,
S.
Navakkode
,
S.Q.Z.
Yeow
,
K.
Krishna-K
,
M.C.
Liang
,
J.H.
Koh
,
R.X.
Wong
,
W.P.
Yu
,
S.
Sajikumar
,
H.
Huang
, and
T.W.
Soong
.
2022
.
Loss of CaV1.3 RNA editing enhances mouse hippocampal plasticity, learning, and memory
.
Proc. Natl. Acad. Sci. USA
.
119
:e2203883119.
Zhang
,
J.F.
,
P.T.
Ellinor
,
R.W.
Aldrich
, and
R.W.
Tsien
.
1994
.
Molecular determinants of voltage-dependent inactivation in calcium channels
.
Nature
.
372
:
97
100
.
Zhang
,
H.
,
Y.
Fu
,
C.
Altier
,
J.
Platzer
,
D.J.
Surmeier
, and
I.
Bezprozvanny
.
2006
.
Ca1.2 and CaV1.3 neuronal L-type calcium channels: Differential targeting and signaling to pCREB
.
Eur. J. Neurosci.
23
:
2297
2310
.
Zhao
,
Y.
,
G.
Huang
,
J.
Wu
,
Q.
Wu
,
S.
Gao
,
Z.
Yan
,
J.
Lei
, and
N.
Yan
.
2019
.
Molecular basis for ligand modulation of a mammalian voltage-gated Ca2+ channel
.
Cell
.
177
:
1495
1506.e12
.
Zhou
,
H.
,
K.
Yu
,
K.L.
McCoy
, and
A.
Lee
.
2005
.
Molecular mechanism for divergent regulation of Cav1.2 Ca2+ channels by calmodulin and Ca2+-binding protein-1
.
J. Biol. Chem.
280
:
29612
29619
.
Zhu
,
L.
,
S.
McDavid
, and
K.P.
Currie
.
2015
.
“Slow” voltage-dependent inactivation of CaV2.2 calcium channels is modulated by the PKC activator Phorbol 12-Myristate 13-Acetate (PMA)
.
PLoS One
.
10
:e0134117.
Zühlke
,
R.D.
,
G.S.
Pitt
,
K.
Deisseroth
,
R.W.
Tsien
, and
H.
Reuter
.
1999
.
Calmodulin supports both inactivation and facilitation of L-type calcium channels
.
Nature
.
399
:
159
162
.

Author notes

Disclosures: The authors declare no competing interests exist.

This article is distributed under the terms as described at https://rupress.org/pages/terms102024/.