Next Article in Journal
Comparison between Three Different Techniques for the Detection of EGFR Mutations in Liquid Biopsies of Patients with Advanced Stage Lung Adenocarcinoma
Next Article in Special Issue
Glutamate-Evoked Ca2+ Responses in the Rat Suprachiasmatic Nucleus: Involvement of Na+/K+-ATPase and Na+/Ca2+-Exchanger
Previous Article in Journal
Short, Lipidated Dendrimeric γ-AApeptides as New Antimicrobial Peptidomimetics
Previous Article in Special Issue
Spatial and Functional Crosstalk between the Mitochondrial Na+-Ca2+ Exchanger NCLX and the Sarcoplasmic Reticulum Ca2+ Pump SERCA in Cardiomyocytes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Regulation of Cardiac Cav1.2 Channels by Calmodulin

1
Department of Physiology, Graduate School of Medical & Dental Sciences, Kagoshima University, Sakura-ga-oka, Kagoshima 890-8544, Japan
2
Department of Pharmaceutical Toxicology, School of Pharmacy, China Medical University, Shenyang 110012, China
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(7), 6409; https://doi.org/10.3390/ijms24076409
Submission received: 25 January 2023 / Revised: 19 March 2023 / Accepted: 23 March 2023 / Published: 29 March 2023
(This article belongs to the Special Issue Calcium Channels and Calcium-Binding Proteins)

Abstract

:
Cav1.2 Ca2+ channels, a type of voltage-gated L-type Ca2+ channel, are ubiquitously expressed, and the predominant Ca2+ channel type, in working cardiac myocytes. Cav1.2 channels are regulated by the direct interactions with calmodulin (CaM), a Ca2+-binding protein that causes Ca2+-dependent facilitation (CDF) and inactivation (CDI). Ca2+-free CaM (apoCaM) also contributes to the regulation of Cav1.2 channels. Furthermore, CaM indirectly affects channel activity by activating CaM-dependent enzymes, such as CaM-dependent protein kinase II and calcineurin (a CaM-dependent protein phosphatase). In this article, we review the recent progress in identifying the role of apoCaM in the channel ‘rundown’ phenomena and related repriming of channels, and CDF, as well as the role of Ca2+/CaM in CDI. In addition, the role of CaM in channel clustering is reviewed.

1. Introduction

Calmodulin (CaM) was discovered in 1970 as a Ca2+ regulator, responsible for nucleotide phosphodiesterase in the brain [1]. Now, CaM is widely known as a ubiquitous eukaryotic Ca2+ sensor protein [2]. A small (16.7-kDa) dumbbell-shaped protein, CaM is ubiquitously distributed in tissues/cells and mediates various Ca2+-dependent signaling. CaM consists of an N-lobe, C-lobe, and linker connecting these two lobes (Figure 1). Each lobe has two EF-hand motifs that allow binding of two Ca2+ ions (four Ca2+ ions as a CaM molecule), thereby controlling various Ca2+-dependent pathways [3,4]. The two lobes have different affinities for Ca2+ and target binding. Affinity of the N-lobe for Ca2+ (Kd ~10 µM) is lower than that of the C-lobe (Kd ~1 µM), meaning that the dynamic range of CaM for Ca2+ spans a wide range, and allows it to transmit intracellular Ca2+ signals to more than 200 targets. CaM changes its conformation from a closed-in Ca2+-free state (apoCaM) to an open-in Ca2+-bound state (Ca2+/CaM) [5], forming the basis of target binding upon Ca2+ concentration ([Ca2+]) changes [6].
Voltage-gated calcium channels (VGCCs) are multimeric proteins (consisting of α1, β, and α2δ subunits; some with γ subunits) found in the membranes of nerve, muscle, and secretory cells [7]. VGCCs are divided into high-voltage-activated L-type (Cav1.1, Cav1.2, Cav1.3, and Cav1.4), P/Q-type (Cav2.1), N-type (Cav2.2), and R-type (Cav2.3) channels, because they are activated by relatively large depolarizations; as well as low-voltage-activated T-type (Cav3.1, Cav3.2, Cav3.3) channels, because they are activated by relatively small depolarizations. Among VGCCs, Cav1.2 is the dominant type in cardiac working muscle, while Cav1.3 and Cav3.2 are dominant in pacemaker (nodal) cells. Ca2+ influx through Cav1.2 channels is regulated by the negative feed-back mechanism, known as Ca2+-dependent inactivation (CDI), as well as the positive feed-back mechanism, known asCa2+-dependent facilitation (CDF) [8]. CaM is thought to play an important role in both CDI and CDF. Furthermore, CaM may have roles in channel trafficking and clustering/multimerization. Here, we review the regulation of cardiac Cav1.2 channels by CaM. Excellent reviews on similar subjects have appeared recently [8,9,10,11,12,13,14,15].

2. Functional Studies

2.1. Rundown of Channels

2.1.1. History

‘Rundown’ of channels is a phenomenon in which the channel activity progressively declines under conditions where the intracellular environment is artificially changed. Rundown has been recognized since the 1970s, during the early period of intracellular recording of Ca2+ channels [16,17,18]. Rundown is thought to be caused by (1) channel dephosphorylation [17,19,20,21,22], (2) a switch-on of the inactivation process due to increased [Ca2+] [23], (3) a loss of activation factors [24,25,26,27,28], and/or (4) channel proteolysis [20,27,29] Thus, the rundown may be produced by multiple causes.

2.1.2. How to Prevent the Rundown

To prevent the rundown of Ca2+ channels and achieve stable recording of channel activity, the ‘perforated’ patch-clamp method was developed [30]. This method does not break the patch membrane in the recording pipette, but instead applies ionophores, such as nystatin, to open small pores in the patch membrane.
Another method to overcome the channel rundown is the replenishment of the factors required for the maintenance of Ca2+ channel activity. The rundown of Cav1.2 channels observed in inside-out patch-clamp mode is rapid and severe, with channel activity progressively decreasing until it disappears within several minutes [16,25,26]. Calpastatin, an inhibitor of Ca2+-activated protease calpain, was the first molecule reported to reverse the rundown [24,27,28]. Although this effect of calpastatin was first thought to be caused by the inhibition of calpain, its acute effect is localized to the non-inhibitory region of calpain [31,32]. Later, CaM (together with ATP, see below) was found to have the same effect [33]. This effect of CaM is independent of protein kinases, such as CaM-dependent protein kinase II (CaMKII), because it is not blocked by the non-specific protein kinase inhibitor K252a, or CaMKII inhibitors KN62, CaMKII peptide 281–309, and autocamtide-2-related inhibitory peptide [33]. Because calpastatin has a small effect compared to CaM and shows competition with CaM for its effect on Ca2+ channel activity, calpastatin is thought to act as a partial agonist for the CaM-binding site of the Cav1.2 channel [34,35].

2.2. Repriming of Channels

2.2.1. CaM Is Crucial for Channel Activity

CaM is thought to be crucial for Cav1.2 channel activity (see above). This view is in line with reports that indicate that CaM associates with both Cav1.2 and Cav2.1 channels to regulate their activity [36,37,38,39]. Mutations of the CaM-binding domain in the carboxy-terminal tail of the Cav1.2 channel α1C subunit disrupted binding of CaM to the channel and resulted in a loss of the regulatory effect of CaM [40,41]. The repriming effect of CaM (with 3 mM ATP at 80 nM Ca2+) on the Cav1.2 channels is concentration-dependent [half-maximal effective concentration (EC50) of ~1 µM] and had a maximum effect of 300–400% of the recordings in the preceding cell-attached patches. If we assume that the ascending phase of the dose-response relationship (Figure 2) simply reflects the binding of apoCaM to the channel, the Kd for CaM binding would be ~1 μM. This value is much higher than that reported for the binding of Ca2+/CaM to fragments of carboxy-terminal regions of the channel (10–100 nM) [38,42,43,44], but is lower than that reported for apoCaM binding to these fragments (~10 μM; [43,44]). Thus, these results suggest that the channel does not tightly interact with apoCaM at resting conditions ([Ca2+] ≈ 80 nM) but may be in dynamic equilibrium with free apoCaM in the cytoplasm near the internal side of the channel. CaM at ~0.3 μM + ATP (3 mM) produced channel activity that is comparable to that observed in the preceding cell-attached mode, suggesting that the concentration of free CaM near Ca2+ channels in myocytes was in this range. This concentration is comparable to that reported for cardiac myocytes, i.e., [CaM]free value of 50–75 nM [45], considering an inhomogeneous distribution of CaM. We speculate that fewer than one-third of the channels are bound by apoCaM under resting conditions. Furthermore, CaM higher than 0.3 μM (+3 mM ATP) produced higher channel activity than that observed in the cell-attached mode. Collectively, these findings suggest a new mechanism of channel modulation, namely, changes in apoCaM concentrations following the release from or the absorption to CaM-binding proteins that modulate Ca2+ channel activity. Interestingly, both the individual N- and C-lobes of CaM can rescue channel activity from rundown, suggesting that CaM may reprime the Cav1.2 channel by binding to either lobe [46]. Notably, half-calcified CaM (Ca2+-bound only to C-lobe) bound to Cav1.2 channels with the C-lobe only (while the N-lobe is free) under resting conditions [see below (Section 3)] [47].
A similar effect of apoCaM was reported for Cav1.3 channels, which did not show a complete rundown in CaM-free conditions but had reduced activity, whereas, the application of apoCaM enhanced Cav1.3 channel activity [48,49]. However, the binding affinity of apoCaM for the Cav1.3 channel is reportedly only ~10 nM [50], meaning Cav1.3 channels are nearly saturated with apoCaM under physiological conditions near the channels (i.e., [CaM]free = 50~500 nM) [45,51].
Figure 2. Effect of CaM in cell-free patches of cardiomyocytes. (A) Concentration-dependent effect of CaM on activity of Cav1.2 channels recorded in inside-out patches of guinea-pig cardiomyocytes at fixed [Ca2+], as indicated by colors on the right side of the graph. Channel activity was normalized to that recorded in the preceding cell-attached mode. Date are represented as the mean ± S.E.M. (n = 6–11) with regression curves with an equation as Amax · ([CaM]/Kdf)nf / {1+([CaM]/Kdf)nf } · 1/{1+([CaM]/Kdi)ni }, where Amax is maximum channel activity, [CaM] is concentration of CaM, Kdf and Kdi are CaM’s dissociation constant for facilitation and inactivation, respectively, and nf and ni are the Hill’s coefficients for facilitation and inactivation, respectively. (B) Relationship between estimated maximum activity (Amax) and [Ca2+]. Data were fitted with Hill’s curve with EC50 ~1 μM (red bar). Reproduced from Han et al. [51] with permission.
Figure 2. Effect of CaM in cell-free patches of cardiomyocytes. (A) Concentration-dependent effect of CaM on activity of Cav1.2 channels recorded in inside-out patches of guinea-pig cardiomyocytes at fixed [Ca2+], as indicated by colors on the right side of the graph. Channel activity was normalized to that recorded in the preceding cell-attached mode. Date are represented as the mean ± S.E.M. (n = 6–11) with regression curves with an equation as Amax · ([CaM]/Kdf)nf / {1+([CaM]/Kdf)nf } · 1/{1+([CaM]/Kdi)ni }, where Amax is maximum channel activity, [CaM] is concentration of CaM, Kdf and Kdi are CaM’s dissociation constant for facilitation and inactivation, respectively, and nf and ni are the Hill’s coefficients for facilitation and inactivation, respectively. (B) Relationship between estimated maximum activity (Amax) and [Ca2+]. Data were fitted with Hill’s curve with EC50 ~1 μM (red bar). Reproduced from Han et al. [51] with permission.
Ijms 24 06409 g002
Numerous proteins interact with Cav1.2 channels [e.g., calpastatin, Ca2+-binding protein 1 (CaBP1), and α-actinin] [52], making them possible candidates for the ‘repriming’ proteins of these channels. In addition, Rad, an RGK protein that interacts with the β-subunit, suppresses Cav1.2 channel activity, e.g., [53]. Future studies are expected to identify the channel-repriming proteins other than CaM.

2.2.2. ATP Is Another Crucial Factor

The repriming effect of CaM requires the presence of ATP because neither CaM nor ATP alone can effectively restore channel activity in inside-out patches [25,26,33]. Because ATP can be partially substituted by AMP-PNP, a nonhydrolyzable ATP analog, a phosphorylation-independent mechanism is suggested (e.g., an ATP-binding site within or close to the channel) [54]. This view is further supported by the experiments that show that the effect of ATP can be partially replaced by other nucleotides with the following order of efficacy: ATP > GTP > UTP > ADP > CTP ≈ AMP [55]. Thus, the presumed nucleotide-binding site in the channel seems to favor a purine rather than pyrimidine base, and a triphosphate rather than a di- or monophosphate group. Notably, high concentrations (> 10 mM) of GTP, UTP, CTP, ADP, or AMP have inhibitory effects on channel activity [55]. Thus, protein phosphorylation is unlikely to be involved in the effect of ATP. A phosphorylation-independent action of ATP was previously reported [27,56]. Exploration for an ATP binding site in the Cav1.2 channel suggests that one occurs in the C-terminal tail of the α1C subunit [57]. Thus, the Cav1.2 channel is reprimed by CaM and ATP for activation upon depolarization. In fact, with CaM and ATP, activity of Cav1.2 channels can be stably recorded in inside-out patch-clamp mode for more than 1 h.

2.3. Slow Rundown of Channels

Contribution of CaMKII

CaM + ATP recovered Cav1.2 channel activity to 93% and 28% of that recorded in the cell-attached configuration, when applied 1 and 5 min after the excision of the patch from the cell, respectively, indicating time-dependent attenuation (called ‘slow rundown’). However, in the presence of 0.33 μM CaMKII, this attenuation was markedly retarded, showing 85% and 75% recovery when applied 1 and 5 min after the formation of the inside-out mode, respectively [58]. Although CaMKII T286D, a constitutively active mutant of CaMKII, effectively prevented a slow rundown similar to wild-type (WT), it failed to induce Ca2+-channel activity (only 2–10% of that recorded in cell-attached mode), thus allowing a ‘fast’ rundown to occur shortly after inside-out patch formation. These results suggest that CaMKII may not play a major role in reversing the fast rundown of channels, and that a large portion of CaMKII sites on the channels responsible for channel activity might already be phosphorylated in the basal condition. In pull-down assays, a GST-fusion protein containing amino acids 1509–1789 of the C-terminal region (CT1, Figure 1C) of guinea pig Cav1.2 treated with CaMKII, showed a higher affinity for CaM compared with CT1 treated with phosphatase. On the basis of this result, we proposed a model in which CaMKII-mediated phosphorylation of the Cav1.2 channel regulates its binding of CaM during the reversal of the rundown [59,60]. Furthermore, it is interesting to note that this effect might be related to the CaMKII-mediated facilitation (see below, Section 2.4.2).
Wang et al. found that CaMKII phosphorylates Thr1603 (guinea pig; Thr1604 in rabbit) within the preIQ domain of the C-terminal tail of the Cav1.2 channel [60]. A mutation of Thr1603 to Asp (T1603D) slowed the channel rundown in inside-out patch mode and abolished the time-dependent attenuation of CaM activity to reverse the rundown. These findings support the view that CaMKII-mediated phosphorylation of the proximal C-terminal (pCT) fragment (CT1) increases its binding with CaM, thereby maintaining the basal activity of the channel, while the dephosphorylation of CT1 showed an opposite effect. A similar effect of protein kinase A (PKA) phosphorylation of CT1 on the binding of CaM to CT1 has also been reported [61] (see also [22]).

2.4. Ca2+-Dependent Facilitation (CDF)

2.4.1. Overview of CDF

The Cav1.2 channel shows both voltage-dependent inactivation (VDI) and Ca2+-dependent inactivation (CDI). In addition, this channel shows both voltage-dependent facilitation (VDF) and Ca2+-dependent facilitation (CDF). CDI and CDF may involve CaM as a Ca2+ sensor [62]. Since the 1970s, the contractile force of the heart has been known to depend on its rate and pattern of stimulation [63]. An increase or decrease of contraction force in response to an increased pacing rate is called the positive or negative Bowditch effect or ‘staircase’ phenomenon, respectively, which are thought to primarily involve a change in Ca2+ release from the sarcoplasmic reticulum (e.g., [64]).
Along with the positive staircase phenomenon of contraction force, a similar phenomenon has been observed in cardiac Ca2+ current induced by a train of repetitive depolarizations, or a strong depolarizing pre-pulse [65,66,67,68,69,70,71,72,73], a process termed Ca2+ current (ICa) “staircase” or “facilitation”. Because the facilitation produced by a strong depolarizing pre-pulse is observed with charge carriers other than Ca2+ (e.g., Na+ or Ba2+), it is thought to be Ca2+-independent and thereby VDF. However, because the facilitation produced by repetitive depolarizations is not observed with Ba2+ as a charge carrier, it is called CDF. Accordingly, enhancement of ICa or IBa following a moderate increase of intracellular [Ca2+] in ventricular myocytes was reported [74,75,76,77,78]. Activity of L-type Ca2+ channels was potentiated by a moderate increase in intracellular [Ca2+] ([Ca2+]i, around 180–400 nM) and suppressed by a further increase in [Ca2+]i (>600 nM) [77,78,79].

2.4.2. CaMKII Mediated CDF

There are reportedly two types of CDF: CaMKII-mediated and CaMKII-independent. CaMKII-mediated CDF was independently reported by three groups [80,81,82]. All three studies reported that the pharmacological inhibition of CaMKII abolished CDF in mammalian cardiomyocytes.
Hudmon et al. reported that CaMKII is tethered to the C-terminus of α1C and mediates CaMKII-dependent CDF [83]. In contrast, Abiria et al. reported that CaMKII is anchored to the β2a subunit [84]. Multiple phosphorylation sites of CaMKII are reported in both α1C and β subunits. Specifically, CaMKII phosphorylates Thr498 (rat) of the β2a subunit, and then Ser439 (rabbit), Ser1517 (do.; corresponding to S1512 in a different isoform of rabbit), Ser1570 (do.), and Thr1603 (guinea pig) of the α1C subunit [60,85,86,87,88]. The cardiac myocytes of a knock-in mouse, in which S1512 (rabbit) and S1570 (do.) of Cav1.2 were mutated to alanine, showed no CDF, suggesting that these residues are phosphorylation sites responsible for CDF [88]. However, the physiological significance of the β2 subunit phosphorylation by CaMKII remains to be clarified [89]. It is of note that the CaMKII- mediated enhancement of ICa plays an important role in the development of isoproterenol-induced cardiac hypertrophy, and that the related gene expressions are significantly reduced by a peptide which inhibits binding of CaMKII to Cav1.2 channels [90,91].
Although CaMKII-mediated CDF of the native Cav1.2 channel is obvious in cardiac myocytes, CDF in in-vitro expression systems is observed only after CDI is masked by a mutation of Cav1.2 [41,42,92]. These findings suggest that unknown factors or additional mechanisms are involved in the manifestation of CDF. Yang et al. reported that cultured rat cardiac myocytes infected with adenoviral vectors harboring Ca2+-insensitive CaM mutant genes had increased ICa and were sensitive to CaMKII inhibitors, suggesting that the overexpression of Ca2+-insensitive CaM can activate CaMKII [93].

2.4.3. CaMKII-Independent CDF

Other studies suggested that CDF is mediated by CaM, but independent of CaMKII [41,42,94]. Zühlke et al. showed that mutation of the IQ domain, a CaM-binding site in the C-terminus of Cav1.2, abolished CDI, and unmasked CDF [41,42] (see also [92]). Our investigation of the effects of CaM and CaMKII inhibitors on CDF and CDI with patch-clamp cell-attached recordings revealed that CDF and CDI were depressed and ultimately abolished by the CaM inhibitor chlorpromazine (1–100 μM) in a concentration-dependent manner. In addition, the CaM inhibitor calmidazolium (1 μM) produced a similar effect. In contrast, CaMKII inhibitors KN-62 (0.1–3 μM) and autocamtide 2-related inhibitory peptide (1 μM) significantly delayed the development of CDF and CDI without depressing either. We hypothesized that CaM is a key molecule for CDF and CDI, and that CaMKII plays a modulatory role in both [79].

2.4.4. Biphasic Effects of CaM

Previous reports indicate the biphasic effects of [Ca2+]i on ICa, specifically increased ICa at low [Ca2+]i and decreased ICa at high [Ca2+]i [77,78]. These effects can be taken as a manifestation of CDF and CDI, suggesting that an interaction of apoCaM, or partially Ca2+-bound CaM, with the channel is necessary for its activation, and that the effect of Ca2+-saturated CaM may be Ca2+-dependent inactivation of the channel. Similar findings have been reported for the Cav1.3 channel [49] and ryanodine receptor (Ca2+-release channel in the sarcoplasmic reticulum) in skeletal muscle [95], whereby CaM is an activator at nanomolar [Ca2+] and an inhibitor at micromolar [Ca2+].
We examined the effects of CaM and Ca2+ on Cav1.2 channels in guinea pig ventricular myocytes in inside-out patch mode, whereby the channel rundown was controlled. At [Ca2+] of 0.1 μM, CaM (0.15–7.0 μM) + ATP (2.4 mM) induced channel activity with a bell-shaped relationship to [CaM], including a peak effect around 2 μM (Figure 2). Similar results were observed at a free [Ca2+] < 0.01 μM, or with the Ca2+-insensitive mutant CaM1234, suggesting that even apoCaM may be able to induce facilitation and inactivation of channel activity. The bell-shaped curve of CaM was shifted to a lower concentration (left) with increasing [Ca2+]. On the basis of these findings, a simple model for CaM- and Ca2+-dependent modulations of channel activity that assumes two CaM-binding sites was proposed (Figure 3). It suggests that both apoCaM and Ca2+/CaM can induce facilitation and inactivation of Cav1.2 channels, and that the basic role of Ca2+ is to accelerate CaM-dependent facilitation and inactivation [51]. We also investigated the concentration- and Ca2+-dependent effects of CaM mutants, CaM12 and CaM34, in which Ca2+ binding to N- and C-lobes is eliminated, respectively, on Cav1.2 channels by inside-out patch-clamping of guinea pig cardiomyocytes. Both CaM12 and CaM34 (0.7–10 µM) applied with 3 mM ATP produced channel activity after the rundown. Concentration–response curves were bell-shaped, similar to wild-type CaM, however, there was no obvious leftward shift of the curves with increasing [Ca2+], as observed for WT CaM, suggesting that both functional lobes of CaM were necessary for the Ca2+-dependent shift [96].
Reported [CaM]free in resting cardiac myocytes (100 nM [Ca2+]) are 50–75 nM, representing only 1% of the total (free + bound to cellular constituents) [CaM] [45,100]. Therefore, CaM-bound Cav1.2 channels at the resting condition (i.e., [CaM]free = 200 nM, considered a high [CaM] in Z-line area [45]) and [Ca2+] = 100 nM) represent approximately 10% of the total number of channels in the cell-free patches [51]. Thus, the potential dynamic range of the CaM-mediated CDF may be larger from a design point of view than considered thus far. Interestingly, the maximum effects of CaM12 and CaM34 were reduced as [Ca2+] increased [96]. A similar phenomenon was observed with WT CaM, suggesting that another Ca2+ sensor might exist in addition to CaM [51].
It is of note that the concentration of the intracellular Mg2+ shows biphasic effects on ICa, facilitation and inactivation, similar to Ca2+ [101]. It may be interesting to explore if the Mg2+ effects are mediated by CaM.

2.4.5. Mutations That Affect CDF

Numerous studies suggest that the IQ domain mutations in full-length channels can eliminate CDI and CDF [41,42,92,102,103]. Thus, it will be interesting to establish how CaM regulates VGCCs other than Cav1.2 through its interactions with their IQ motifs (reviewed elsewhere [11,15,104,105,106]. The Cav1.2 IQ domain residues F1618, Y1619, and F1622 (rabbit) are involved in the interaction with the N-lobe of Ca2+/CaM, and triple-alanine (TripleA) mutation of these sites eliminated CaM-mediated CDF [83,92], suggesting a hydrophobic interaction between the N-lobe of CaM and the IQ domain that is important for CDF. Kim et al. reported that the mutations in the preIQ-A (L1562A/F1563A, human α1C, 77) and the preIQ-C (W1593A and W1593E) together with I1624A (which unmasks CDF) reduced CDF, suggesting an important role for the Ca2+/C-lobe-preIQ interaction in CDF [107].
To date, there is no established CDF conformation of Cav1.2 and CaM. It is possible that multiple mechanisms and thus multiple conformations exist for CaM-dependent CDF of Cav1.2 channels.

2.4.6. Physiological Significance of CDF

As discussed above, CDF seems to involve at least two different mechanisms: one mediated by CaMKII phosphorylation, and one mediated directly by CaM. The latter CDF occurs rapidly in response to local [Ca2+]i on a beat-to-beat basis, making Ca2+ influx “all-or-nothing” to ensure the reliability of excitation-contraction coupling. CDF mediated by CaMKII phosphorylation occurs more slowly (over several beats) and seems to have a different role. One study has proposed that the facilitatory mechanism may partly offset reduced Ca2+ channel availability at high heart rates (caused by increased CDI), contributing to the improvement of cardiac performance during exercise [108].

2.5. Ca2+-Dependent Inactivation (CDI)

2.5.1. Overview of CDI

CDI was first demonstrated by Brehm and Eckert (1978) in voltage clamp recordings of the paramecium [109]. Eckert and collaborators also showed that the buffering of intracellular Ca2+ reduced the inactivation of ICa, confirming the necessity for an increase of [Ca2+]i for CDI [109,110,111]. CDI of Ca2+ channels was subsequently reported in various eukaryotic cell types across kingdoms, including placozoans, poriferans, ciliates, and vertebrates [112,113]. Among Cav channels in vertebrates, Cav1.2, Cav1.3, and Cav2.1 (P/Q type) channels show definitive CDI, while Cav1.1 shows modest CDI, Cav2.2 (N type) and Cav2.3 (R type) channels show tiny or undetectable CDI, and Cav1.4 and Cav3 (low-voltage activated type) channels show vestigial CDI [114,115].
The involvement of CaM in CDI is now widely accepted. Birnbaumer and colleagues found that the binding of CaM to the IQ domain of the C-terminus of the α1C subunit triggers CDI [116,117]. Reuter and colleagues [103,118] also reported that CDI of the Cav1.2 channel depends on the IQ motif of the C-terminus of the α1C subunit. Yue and collaborators also reported the involvement of CaM in CDI [102].

2.5.2. Dephosphorylation vs. Ca2+/CaM

A mechanism other than the direct interaction of Ca2+/CaM has been suggested. Specifically, the Ca2+-dependent protein phosphatase calcineurin (CaN) can also accelerate the inactivation of the L-type Ca2+ currents during depolarization. Because CaN acts in a Ca2+/CaM-dependent manner, CaN-mediated dephosphorylation provides another molecular mechanism for CDI of Cav1.2 channels [119]. This hypothesis predicts attenuation or slowing of CDI by PKA phosphorylation, a phenomenon described for L-type currents in some neurons [119,120], but not Cav1.2-mediated currents in cardiomyocytes [12,121,122]. Because Cav1.2 channels are markedly upmodulated by PKA, the counteraction by responsible phosphatases should also be robust. If the time course of this action is slow, then the phosphatase effect would manifest as a down-modulation or slow CDI.

2.5.3. Mutations That Affect CDI

De Leon et al. reported that pCT regions containing EF-hand, preIQ, and IQ domains could transfer CDI from α1C to α1E, which had little CDI, suggesting that CDI was mediated by these domains [123]. Zühlke et al. explored important regions for CDI by investigating truncation mutants in the pCT of Cav1.2 channels [103]. They found three regions [i.e., the EF-hand domain, two hydrophilic residues (asparagine and glutamic acid) in the A region of preIQ, and IQ] that were important for CDI. Soldatov et al. revealed that IKTEG1572-76 in the preIQ-A region (human α1C,77) and LLDQV1600-04 in the preIQ-C region are important for CDI [124]. Qin et al. reported that the mutation of six residues in the IQ domain eliminated CDI [117], and Peterson found that a mutation replacing VVTL in the EF-hand domain (rabbit α1C) with the corresponding MYEM of α1E, which shows no CDI, eliminated CDI [125]. Furthermore, V1548Y (rabbit) mutation of the same region was enough to significantly reduce CDI. The authors suggested that this region might play a ‘transducer’ role rather than act as a sensor of CDI (see below 4.4. Downstream mechanism of CDI).
Since then, a number of mutation experiments around this area have been reported. Kim et al. reported that the mutation of TLF in the preIQ-A region abolished CDI [126]. It was confirmed that T1561A and L1565A in the preIQ-A (human α1C,77) reduced CDI [107], but not L1585A/L1588A in the dimer interface [127] nor W1593E in the preIQ-C affected CDI, thus confirming the importance of the preIQ-A region. In the same report, a W1593A mutation significantly reduced CDI [107], suggesting that the preIQ-C region was also important. Zühlke et al. reported that a I1624E (human α1C,77) mutation reduced the affinity of the IQ for CaM ~100-fold and eliminated CDI [42]. Erickson et al. reported that a IQ·FRK1654-60/5A (rabbit) mutation resulted in almost no CDI when WT CaM was co-expressed [128]. It has been reported that an I/E mutation in the IQ domain of Cav1.2 of transgenic mice abolished the regulation of CDI and CDF similar to findings of in vitro expression systems [129,130].
The importance of the pCT domain of the Cav1.2 channel for CDI is thus unequivocally documented. However, Dick et al. reported that the NT region also plays a role in CDI by bridging CaM between the NT and CT domains (NT-CT bridging model) [98]. The N-lobe of Ca2+/CaM interacts with a region named the N-terminal spatial Ca2+ transforming element (NSCaTE: 81SWQAAIDAARQAK93, rabbit), in which W82A, I86A, and R90A mutations significantly reduced CDI. Accordingly, whether the CT-restricted model or the NT-CT bridging model of CDI operates in intact myocytes remains an open question. It is possible that both mechanisms work simultaneously in myocytes.
Mutations of CaM have been also reported to affect CDI (reviewed by Crotti et al. [131]). Among them, mutations of C-lobe are most frequently found, e.g., E141G [132].

2.5.4. Interactions of Domains within C-Terminal Tail of Cav1.2

Hulme et al. reported that a domain immediately after IQ (termed the proximal C-terminal regulatory domain, PCRD) and a domain near the C-terminus (termed the distal C-terminal regulatory domain, DCRD) bind to each other, leading to a decrease in open probability [121]. Minobe et al. suggested a possible molecular mechanism by which the PCRD-DCRD interaction suppresses channel activity to disrupt the facilitatory effect of apoCaM and/or ATP on channel activity [133]. Crump et al. investigated the effects of the distal C-terminus of Cav1.2 (a.a. 1822–2171 of rabbit) on the activity of the channels and found that it inhibited IBa,L but not ICa,L, and that its blockade was antagonized by Ca2+/CaM [134]. From these findings, they proposed a model in which the distal region functions as a reverse use-dependent inhibitor of the Cav1.2 channel.
Lyu et al. reported a new interaction between the proximal (containing preIQ and IQ regions) and distal (distal of DCRD, a.a. 2116–2169 of guinea pig) regions of Cav1.2. Because binding is inhibited by Ca2+/CaM, the new distal interaction domain was named the “CaM-competitive domain” (CCD) [135]. Electrophysiological experiments showed that a CCD peptide inhibited, while a preIQ-IQ peptide facilitated, Cav1.2 channel activity in inside-out patches of guinea pig ventricular myocytes. These results suggest that the distal CCD inhibits the Cav1.2 channel by modulating its CaM-binding properties. The CCD overlaps with the distal ends of the C-terminal modulator (CTM) of Cav1.3 [136,137,138] and the inhibitor of CDI region of Cav1.4 [50,139,140], both of which are known to modulate CDI. The CTM of Cav1.3 has also been shown to compete with CaM for binding to the pCT region [50,138,141].

2.5.5. C-Terminal Modulator of Cav1.3 and Cav1.4

Striessnig and colleagues found that an auto-modulatory domain within the CT of Cav1.3 and Cav1.4 precisely controls the gating properties of these channels [136,137,140,142,143]. This CTM, which consists of two α-helices in the PCRD and DCRD C-terminus, strongly reduces the CDI of channels and promotes their activation at more negative voltages [142]. Thus, the C-terminal splice variants of Cav1.3 show different CDI and activation properties [144]. Similar regulation mediated by the CTM has also been reported [139,145].

2.5.6. Lobe Specific Effects of CaM

The N- and C-lobes of CaM appear to have distinct roles in CDF and CDI. For example, Ca2+-bound N-lobe produces CDI of Cav2.1 (P/Q type) channels, while Ca2+-bound C-lobe produces CDF [40]. For the Cav1.2 channel, CDI has only been attributed to the C-lobe of CaM [102,146,147]. However, Dick et al. observed CDI in HEK-293 cells expressing Cav1.2 and CaM34, which was abolished by the mutation of Trp82 to alanine in the N-terminal NSCaTE domain of the channel [98]. They proposed a model in which the C-lobe of CaM anchors to the C-terminus of Cav1.2, while the N-lobe of CaM binds to the N-terminus of the channel, bridging the two terminals to produce CDI. Ben-Johny and Yue summarized roles of the N- and C-lobes of CaM in sensing local or global [Ca2+] for Cav1 and Cav2 channels [11]. For the Cav1.2 channel, both lobes of CaM sense local Ca2+. However, it is difficult to believe that each lobe of CaM senses distinct local and global [Ca2+] because the distance between the two lobes is only ~1 nm. Thus, it may be better to consider these lobes in the temporal domain rather than in the spatial domain, with both decoding [Ca2+] in the same place with distinct temporal characteristics [148]; i.e., the C-lobe senses a rapid change of [Ca2+] with fast kinetics, while the N-lobe senses time-averaged [Ca2+] with slow kinetics [149].
Limpitikul et al. reported a mathematical model of CDI with decomposed effects of individual N- and C-lobes. This bi-lobal model of CDI indicates high cooperativity of both lobes in CDI [150]. Based on a study using nuclear magnetic resonance (NMR), it is proposed that the C-lobe of CaM selectively binds to its target (i.e., binds first to its target) [151], while the N-lobe binds afterwards, through an induced-fit or coupled conformational selection mechanism initiated by the C-lobe of CaM [152]. The idea that the C-lobe of CaM contributes to selectivity while the N-lobe contributes to flexibility is supported by recent molecular dynamics simulations [153].

2.5.7. CaM-Dependent Inactivation

Han et al. found that even apoCaM and the Ca2+-insensitive mutant CaM1234 can induce inactivation of Cav1.2 channels at high concentrations, referred to as “CaM-dependent inhibition” [51]. These results imply that a change in [CaM] alone could affect Cav1.2 channel activity. In neurons, [CaM] is regulated by small neuronal IQ-motif proteins in response to cellular signals [154]. Recently, such proteins were also found in muscle cells [155]. Moreover, the modulatory effects of the growth-associated protein 43 (GAP43, neuromodulin) on ryanodine receptors and Cav1.1 channels have been reported [156]. Thus, it will be interesting to investigate whether such effects occur for Cav1.2 in cardiac myocytes.

2.5.8. Additional Ca2+-Sensing Site

At [Ca2+] higher than 0.5 μM, the repriming or facilitatory effect of CaM was attenuated with an apparent half-maximal inhibitory concentration of 0.9 μM, as if inactivation would have taken place. This result was not due to an inhibitory effect of Ca2+/CaM because this effect was also observed for the Ca2+-insensitive mutant CaM1234. Therefore, it is suggested that an ‘additional’ Ca2+ sensor for CDI exists other than CaM. Indeed, it is possible that Ca2+ interacts directly with the channels to promote a low-sensitivity CDI [51].

2.5.9. CaBP1-Mediated Regulation

Numerous proteins interfere with CaM binding to pCT, including CaBPs, RGK proteins [157], α-actinin [44,158], and calpastatin [35]. Among these proteins, CaBP1 is most extensively studied. CaBP1 is a member of CaBP family, and highly expressed in the brain and retina [159]. CaBP1 suppresses CDI of Cav1.2 and brings about CDF [126,160]. CaBP1 binds to the IQ domain of Cav1.2 in a Ca2+-dependent manner (higher affinity with Ca2+ than without Ca2+) and competitively inhibits CaM binding to this domain [126,160,161,162,163,164]. However, CaBP1 also reportedly binds to the NT region, whose presence is required for a CaBP1-mediated inhibition of CDI [160].
CaBP1 produces a CaMKII-independent CDF by a different mechanism than a CaM-mediated CDF. Residues of F1618, Y1619, and F1622 in the IQ domain of Cav1.2 (human α1C,77) are involved in the interaction between the N-lobe of Ca2+/CaM and the IQ domain and play a role in CaM-mediated CDF of Cav1.2 [162,163]. A tripleA mutant of these residues eliminated CaM-mediated CDF [92] but not CaBP1-mediated CDF or the inhibition of CDI. Thus, Cav1.2 CDF mediated by CaM and CaBP1 have been suggested to use different molecular mechanisms.
CaM-mediated CDF of Cav1.2 was unmasked by a mutation of the IQ domain, I1624A [41,42,92]. However, co-expression of I1624A mutant channel with CaBP1 produced CDF with a magnitude indistinguishable from that of WT Cav1.2 with CaBP1, confirming that mechanisms of CDF mediated by CaM and CaBP1 are different. Findeisen et al. suggested that the N-lobe and interlobe linker of CaBP1 are important for inhibition of CDI and induction of CDF in Cav1.2 channels [161]. Findeisen et al. proposed a model in which the C-lobe of CaBP1 anchors to the IQ domain and the N-lobe of CaBP1 binds to the NT of Cav1.2 as a modulator [161,162], the latter being previously suggested [160]. Further details on regulation of Cav1.2 channels by CaBPs are described elsewhere [114].
The Cav1.3 channels in auditory inner hair cells show weak or no CDI, in contrast to the same channels in other tissues. Co-expression of CaBPs eliminates CDI of the Cav1.3 channel [165,166].

3. Molecular Biological Studies

3.1. Ca2+/CaM Binding Site

The binding of CaM to the IQ segment was first reported in the late 1990s [41,102,117]. This interaction is Ca2+-dependent with a 1:1 stoichiometry and occurs at Ca2+ concentrations above 100 nM. However, competition experiments with mutant CaMs defective in Ca2+ binding suggest that CaM is constitutively tethered to the channel [102]. The IQ motif and CaM are involved in both the facilitation and the inactivation of Cav1.2 [42]. Subsequently, several regions of the channel, including the NT, I-II loop, and preIQ (AC or CB domain) domains were identified [39,103,124,125,167,168,169].
Mouton et al. determined the Kd of the preIQ (CB domain) and IQ domains of Ca2+/CaM binding to be 12 and 40 nM, respectively [38]. They suggested that the ternary complex formed by preIQ, IQ, and CaM forms the switch leading to CDI.

3.2. ApoCaM Pre-Association Site

Pre-association of apoCaM with Cav1.2 channels is proposed to function as a Ca2+ sensor for their Ca2+-dependent regulation [41,102,117]. Several regions of this channel, including the NT, I-II loop, preIQ (CB domain or AC region), and IQ domains, have been suggested as apoCaM-binding sites [39,103,124,125,167,168,169]. Among these, a region that includes both the preIQ and IQ domains in the C-terminal tail seem to be widely accepted as the pre-association site for CaM [41,102,117]. Furthermore, the IQ domain is required for both CDF and CDI [41,102]. Although the dual role of CaM in regulating CDF and CDI suggests multiple binding sites for CaM in the channel, a single CaM-binding site with multiple conformations depending on the Ca2+-binding state of CaM has also been suggested. Furthermore, phosphorylation mediated by Ca2+/CaM-dependent protein kinase II (CaMKII) may play a crucial role in CDF [170]. However, the precise molecular mechanisms for CDF and CDI remain to be clarified.
It is natural to postulate that Cav1.2 channels, the activity of which is modulated by CaM-sensed Ca2+, pre-associate apoCaM at low [Ca2+]i. This is the case for voltage-gated Nav1.2/1.5 channels, P/Q-type Ca2+ channels, the ryanodine receptor, Ca2+-activated K+ channels (SK and IK type), and the ether-a-go-go-type K+ channel (for review see [105,171,172,173,174]). Although the binding of Ca2+/CaM to Ca2+ channels, including the IQ domain, has been consistently demonstrated [41,102,117], the binding of apoCaM to channels was controversial in the early stage [39,102,117,167,168,175]. Later, Erickson et al. (2001; 2003) used fluorescence resonance energy transfer (FRET) experiments to provide clear evidence that apoCaM tethers to Cav1.2 channels near the IQ domain (< 100 Å) [128,176]. They raised two points for possible features of the pre-association: (1) pre-association of CaM could be coordinated by multiple segments in the proximal CT region, and (2) pre-association could require resting [Ca2+] for proper folding of a channel binding pocket for apoCaM [39,175]. Tang et al. reported a similar conclusion, namely that a peptide containing preIQ-C and IQ domains had the highest affinity to apoCaM [169].
Disruption of the binding site in the IQ domain for the N-lobe of Ca2+/CaM by simultaneous mutation of three N-lobe aromatic anchor residues (Phe1618, Tyr1619, and Phe1622 of human) to alanine eliminated CDF, suggesting that apoCaM tethers to the IQ domain [92]. This finding is supported by NMR showing that the C-lobe of apoCaM interacts with an IQ peptide [177]. Black et al. investigated the Ca2+ binding properties of CaM bound to IQ peptides of Cav1.2 and other types of Ca2+ channels [178]. They found that the binding of CaM to the IQ peptides increases the Ca2+ affinity of both CaM lobes. Evans et al. investigated the binding of apoCaM to the preIQ-A (a.a. 1588–1609 of rabbit), preIQ-C (a.a. 1614–1635), IQ (a.a. 1644–1670), and IQ’ (a.a. 1650–1675) regions [43]. Although Ca2+/CaM bound to all four peptides with low Kd values (≤2 nM), apoCaM bound to all four peptides with high Kd values (>10 μM). Among them, the IQ displayed a relatively high affinity (Kd ≈ 14 µM). These results suggest that apoCaM pre-associates with the IQ domain, whose N-terminal side (a.a. 1644–1649) is important for binding.
Further evidence supporting the IQ region as an apoCaM-binding site is that the calpastatin L domain (CSL), which partially reprimes channel activity, and is thus thought to bind the apoCaM-binding site, competes with CaM for binding to the IQ domain [34]. Minobe et al. investigated the effect of competition between CSL and CaM on channel activity and found that CSL suppressed the channel-activating effect of CaM in a reversible and concentration-dependent manner [35]. Furthermore, CSL suppressed the binding of CaM to IQ in a competitive manner in a GST pull-down assay, however, it was later reported that preIQ also binds CSL [179]. Turner et al. reported the NMR structure of apoCaM bound to the IQ domain of Cav1.2. The C-lobe of CaM bound to Ile1654, Phe1658, and Lys1662 of the IQ domain (rat), while the N-lobe did not bind to IQ [44]. Thus, the possibility that the preIQ domain includes an apoCaM-binding site is not ruled out.
Black and Persechini investigated binding conformations between CaM and an IQ peptide with fluorescence experiments. When the C-lobe of CaM was Ca2+-free, it directed the N-lobe (probably Ca2+-free) to bind to a site within the IQ domain [180]. In contrast, when the C-lobe was Ca2+-bound, it directed the N-lobe (probably Ca2+-bound) to bind to a site upstream of the IQ domain, or a distant site. It will be interesting to determine how these conformations relate to those associated with CDF and CDI.
Very recently, Bartels et al. reported that half-calcified CaM (Ca2+/C-lobe and apoN-lobe) promoted basal activity of the Cav1.2 channel [47] (see also [15]). This finding is based on their experimental observations that a CaM34 mutant, in which Ca2+-binding ability of the C-lobe is disabled, abolished the increasing effect on the open probability of Cav1.2 channels. They estimated that ~50% of Cav1.2 channels were bound to half-calcified CaM under basal conditions. In the future, it will be interesting to compare the effects of apoCaM and half-calcified CaM on Cav1.2 channel activity.
Notably, a mutation of the central linker of CaM, which connects the N- and C-lobes, markedly reduced apoCaM binding to preIQ and IQ domains, suggesting that the linker of CaM contributes to the binding of apoCaM to Cav1.2 [181].

3.3. Models of the Conformation of CDF and CDI

Zühlke et al. proposed a model for CDF and CDI in which apoCaM is bound to a site near the IQ domain at resting [Ca2+]i and Ca2+ binds to the C-lobe of CaM at moderately increased [Ca2+]i, resulting in a conformational change that favors CDF [42]. At higher [Ca2+]i, fully-saturated Ca2+/CaM leads to further structural changes of the CaM-channel complex that favors CDI. Pitt et al. extended their hypothesis for Ca2+-dependent inactivation of L-type channels, stating that CaM is bound to preIQ-A and -C regions at resting Ca2+ levels, but upon the elevation of [Ca2+]i, the C-lobe of CaM shifts and binds to the IQ domain while the N-lobe remains anchored on the preIQ-A domain [39]. This hypothesis suggests that the binding of a single CaM to sites in the CT domain of Cav1.2 causes conformational changes of the channel depending on the Ca2+-bound condition of CaM. Similar models with different details were proposed by Tang et al. (2003) [169] and Erickson et al. (2003) [128].
Kim et al. proposed a model in which three regions of Cav1.2 (i.e., CaM-binding IQ domain, EF-hand motif, and I-II intracellular loop) are involved [126]. Specifically, the CaM-binding IQ domain of α1C transmits the CDI signal via the α1C EF hand motif to the putative pore-occluder, the α1C I-II intracellular linker [62,126]. Ile1654 is important for this transmission of the CDI signal from IQ to EF domains because mutation of I1654A abolished CDI [97]. CDF might be produced by burying the Ile of the IQ motif and preventing its interaction with the EF hand [62]. Ames et al. [15] and Yadav et al. [99] proposed a similar model of CDI in which the C-lobe of apoCaM, or half-calcified CaM (Ca2+/C-lobe and apoN-lobe), interacts with the IQ domain while the N-lobe is free, allowing IQ to also interact with the EF-hand domain in the pCT. This conformation prevents interactions between the EF-hand domain and the inactivation gate in the III-IV linker of Cav1.2 (instead of I-II linker, see above), maintaining the channel in an activatable state. When [Ca2+]i is elevated, the N-lobe of fully calcified Ca2+/CaM interacts with the IQ domain, making the EF-hand domain release IQ and instead engage with the III-IV linker, leading to inactivation of the channel. Thus, this model suggests that Ca2+ binding to the N-lobe of CaM to IQ is critical to trigger CDI (IQ switch model).
We proposed the “Two CaM” hypothesis in which one CaM binds to the first high-affinity (facilitation) site of the channel to promote repriming and CDF, while the other CaM binds to the second low-affinity (inactivation) site of the channel to promote CDI [51]. This model explains why even apoCaM can produce channel inactivation when a high enough concentration is applied. Asmara et al. [182] and Minobe et al. [35] extended this hypothesis by suggesting that apoCaM dynamically binds to the IQ or preIQ-IQ regions to reprime the channel for opening upon depolarization. When [Ca2+]i is elevated, the first CaM binds with Ca2+ and transforms the channel to a state that favors CDF, whereas the second Ca2+/CaM binds to the preIQ domain to complete CDI.
For Cav2.1 channels, Catterall and Few proposed a CDF and CDI model in which local rises in [Ca2+]i lead to Ca2+ binding to the C-lobe of CaM in a pre-associated site, causing translocation to the IQ domain and CDF [183]. Further rises in [Ca2+]i causes CaM to be fully liganded and interact with both IQ and CBD (=preIQ) domains to produce CDI.
A quite different model involving the N-termini of Cav1 channels has also been proposed [98]. In this model, the N-lobe of Ca2+/CaM binds to a N-terminal region (NSCaTE) in a Ca2+-dependent manner to bridge the N-terminus (N-lobe of Ca2+/CaM bound) and C-terminus (C-lobe of Ca2+/CaM bound) of Cav1.2 and Cav1.3 channels, thereby accelerating their closure [98,148]. This model is supported by other groups, e.g., Liu & Vogel reported the NMR solution structure of a Ca2+/CaM-N lobe-NSCaTE peptide [184]. The authors pointed out that Ca2+/CaM acts almost like an “adaptor” protein to bridge the N- and C-terminal domains of the channel. Simms and Zamponi proposed a similar model in which apoCaM tethers to the preIQ/C (PCI-2)-IQ region and, upon [Ca2+] elevation, the C-lobe of Ca2+/CaM moves to a more proximal C-terminal region (PCI-1), and the N-lobe binds to the N-terminus of the Cav1.2 channel, thereby producing CDI [106].
Although CaM binds to the N-terminus of Cav1.2, a mutation in NSCaTE of Cav1.2 (Trp82, Ile86, and Rrg90 to TripleA) abolishing the binding of CaM only slightly weakened the CDI [185]. The authors subsequently reported that the NT and pCT domains directly interact, regardless of whether pCT binds with apoCaM, and Ca2+/CaM reduces this interaction. On the basis of these findings, they claim that dissociation of the NT-CT interaction leads to CDI, proposing a model in which the closed state of the channel configures the NT-CT interaction with apoCaM bound to CT; this interaction is dissociated in the open state until a second Ca2+/CaM binds to NT during CDI [186]. Furthermore, although preIQ and IQ motifs are reportedly conserved in all L-type channels of metazoans, the NSCaTE motif is missing in some arthropod species and lower animals, such as platyhelminths and cnidarians. Regardless, these species show clear CDI, suggesting that CDI mediated by NSCaTE is an optional mechanism developed later in the evolutionary process [113,187].
Kawaji et al. [188] and Minobe et al. [189] carried out similar experiments and found that even with CaM1234-linked Cav1.2 channels, externally applied Ca2+/CaM could induce CDI and that NT-deleted channels failed to support CDI. From these results, they suggest that two types of CDI might coexist, one produced by two CaMs at CT, and the other produced by one CaM with NT-CT bridging.
For Cav1.3 channels, Ben-Johny et al. proposed a model of CDI based on systematic Ala mutagenesis of the entire C-terminal tail of the channel [190]. In this model, apoCaM pre-associates with the IQ domain through the C-lobe, and a hot-spot between the EF-hand and preIQ domain by the N-lobe. Upon [Ca2+] elevation, the N-lobe of Ca2+/CaM departs to the N-terminal NSCaTE, while the C-lobe of Ca2+/CaM moves to the EF-preIQ region. This model is similar to the one that was proposed by Simms and Zamponi [106], however, small differences exist between these two models, including the N-lobe of apoCaM binding to a region upstream of the preIQ domain in Cav1.3, but to preIQ in Cav1.2. Although the sequence between the EF-hand domain to preIQ (1555QPPLG-RTALR1601, rabbit Cav1.2) shows high homology between Cav1.2 and Cav1.3, CaM binding to this region of Cav1.2 has yet to be verified.
Overall, models for CDF and CDI remain controversial. Specifically, the contributions of one or two CaMs, and whether regulation of CDI is limited to CT or involves both NT and CT remain to be elucidated. Further studies are needed to reach unequivocal conclusions.

3.4. Downstream Mechanism of CDI

The molecular basis of mechanisms downstream of CDI has also been investigated. There are three hypotheses for the “endpoint” mechanism: (1) the hinged-lid theory states that a cytoplasmic inactivation particle plugs the pore, analogous to Nav channels and Shaker type Kv channels [191]; (2) collapse of the selectivity filter (SF) occurs, similar to C-type (or P-type) inactivation of K channels; and (3) allosteric inhibition of an intracellular activation gate occurs by a mechanism distinct from that of VDI [75,97,161,192,193,194,195,196]. Tadross et al. investigated the molecular endpoints of CDI by systematically mutating the intracellular S6 segments of all repeats (I to IV) of the Cav1.3 channel [197,198]. They found that allosteric inhibition of the intracellular activation gate is most likely involved in CDI, while a hinged-lid with a shield (that repels lid closure) mechanism is implicated in VDI; this idea is supported by others in Cav1.2 as well as Cav1.3 [199,200,201,202]. Although Abderemane-Ali et al. [203] suggest that the N-terminal cytoplasmic region [98,148,167,190] and Cavβ/I-II loop complex [97,161,200] affect CDI, how conformational changes in these cytoplasmic regions lead to CDI remains unsolved [97,161,196,197,204,205]. Interestingly, Guggenheimer et al. (2016) have proposed a hypothesis that Ca2+/CaM-induced changes in the NT-CT interaction may, in part, accelerate CDI of CaV1.2 [186].
Other studies reported a close link between ion selectivity and CDI, based on the properties of CDI and SF being simultaneously affected by the mutations of SF or changes in extracellular [Ca2+] [193,206], and the observation that Gd3+ blockade of SF appears mutually exclusive with inactivation [204]. The existence of an SF gate and inner gate formed by the S6 helices [186,197,198,199,207,208] suggests that VGCCs have two gates to regulate their activity. Thus, the molecular endpoint of CDI remains an open question. In addition, whether CDI and VDI share common pathways [97,161,195], or act by independent mechanisms [196,198], needs to be addressed.

3.5. CaMKII Phosphorylation Sites

Phosphorylation of the β2a subunit responsible for CaMKII-mediated facilitation of Cav1.2 channels has been observed [87], suggesting that this phosphorylation modulates the interaction between the β subunit and the C-terminal tail of the α1 subunit, thereby manifesting CDF [87,209,210]. However, a site in the I-II loop (rabbit Ser439) and two sites in the CT1 domain (rabbit Ser1517 and 1575) are functionally phosphorylated by CaMKII [85,86,88], where the roles of these sites in Ca2+-dependent regulation are currently unclear.
We investigated a hypothetical CaMKII phosphorylation site on Cav1.2 that contributes to channel regulation and found that CaMKII phosphorylates Thr1603 (guinea pig, Thr1604 in rabbit) within the preIQ domain of the C-terminal tail of the guinea pig Cav1.2 channel. Mutation of T1603D slowed channel rundown in inside-out patch mode and abolished the time-dependency of CaM effects to reverse rundown. We also found that CaMKII-mediated phosphorylation of the pCT fragment (CT1) increased, while dephosphorylation of CT1 decreased its binding with CaM. These findings suggest that CaMKII regulates the binding of CaM to the Cav1.2 channel, thereby maintaining its basal activity [60].

3.6. Two CaM Hypothesis

Ca2+/CaM binds to multiple peptides derived from α1C of the Cav1.2 channel [98,160,167,169,190], in order of decreasing affinity, IQ > preIQ > I-II loop > N-terminal peptide [35,160,182]. Early studies reported that the binding of CaM to IQ and CB (preIQ) domains is competitive, suggesting only one CaM can bind to the C-terminus [168,169,211,212]. Mori et al. reported that one CaM per channel was sufficient to produce CDI for Cav1.2 channels covalently fused with a single CaM molecule [211]. Furthermore, live-cell FRET analysis of interactions between CaM and Cav1.2 channels suggest a 1:1 CaM/channel ratio [213].
A more recent study showed that a peptide containing both preIQ and IQ domains (Leu1599–Leu1668) bound with approximately 2 molecules of CaM per peptide, supporting the hypothesis that two CaM molecules can simultaneously bind to the C-terminus of the Cav1.2 channel, and modulate its facilitatory and inhibitory activities [182]. Two CaM molecules were simultaneously bound to the C-terminus of the Cav1.2 channel in crystallographic studies [107,127,214], and the Two CaM model is supported by other studies [35,43,51,186,215,216]. Evans et al. investigated the binding of CaM to preIQ and IQ domains by monitoring the increases in the fluorescence anisotropy of fluorescein-conjugated C-terminal peptides of Cav1.2 [43]. They found that Ca2+/CaM bound to both preIQ and IQ domains favorably (Kd ≤ 2 nM), proposing a CDI model in which the first CaM binds to IQ and the second CaM binds to preIQ by making an A-C hairpin formation that bridges the A and C regions. Notably, both preIQ-A and C regions are reportedly important for CDI (see ‘Mutations that affect CDI’ above) [107].
At high [Ca2+], 2 molecules of Ca2+/CaM bind to Cav1.2 channels [217]. Chakouri examined whether single or multiple CaMs are involved in Ca2+-dependent regulation of VGCCs using two CaM or Ca2+-insensitive CaM mutants (CaM1234) tethered to the α and/or β subunit [218]. They found that VGCCs respond to signaling from the α-subunit-tethered CaM and concluded that Ca2+ regulation depends on the pre-association of a single CaM with the α-subunit. However, Minobe et al. found that when the Gly linker connecting CaM with the Cav1.2 channel was short (<12 a.a.), the binding of CaM was so strong that externally applied secondary CaM could not produce its effect on Cav1.2 [189]. Thus, it is possible that the use of a longer Gly linker could produce a different conclusion.

3.7. ATP-Binding Site (s)

Feng et al. examined ATP binding to the Cav1.2 channel using a photoaffinity method [57]. A photoactive biotin-conjugated ATP [8-azido-2′/3′-O-(2-aminoethyl-carmamoyl)-ATP-biotin] labeled both the proximal C-terminal region (a.a. 1509–1789 of guinea pig) and N-terminal region (a.a. 6–140) of the Cav1.2 channel. Interestingly, ATP labeling of the Cav1.2 channel was enhanced by the presence of CaM (~4 fold with CaM: Cav1.2 = 1:1 compared with that in the absence of CaM). Thus, it is possible that CaM and ATP synergistically affect Cav1.2 channel activity.

4. Structural Studies

Van Petegem et al. reported the crystallographic structure of a Ca2+/CaM-bound IQ domain (a.a. 1611–1644 of human α1C,77) showing the complex with a parallel orientation and binding governed mainly by hydrophobic interactions. Phe1618, Tyr1619, and Phe1622 interact with the Ca2+/CaM N-lobe, whereas Tyr1627, Phe1628, and Phe 1631 interact with the Ca2+/CaM C-lobe [92]. Isothermal titration calorimetry analysis revealed two Ca2+/N-lobe binding sites (Kd ≈ 50 nM and Kd ≈ 20 μM) and a single high-affinity Ca2+/C-lobe site (Kd ≈ 2 nM) in the IQ domain. The high-affinity site of the N-lobe overlaps with that of the C-lobe. Fallon et al. reported a similar crystal structure analysis for a complex of Ca2+/CaM and IQ peptide (a.a. 1665–1685 of human Cav1.2, Q13936) [219]. Because the Ile1672 residue (thought to form an intramolecular interaction that induces CDI) was buried, they suggested that this form might be associated with CDF.
Fallon et al. extended the CaM-binding peptide to a preIQ-IQ peptide and explored its CaM-binding activity [127]. They found that four CaM molecules bound to two preIQ-IQ peptides, raising the possibility of Cav1.2 channel dimerization (Figure 4A). Kim et al., however, reported that the channel formed monomers but not dimers in Xenopus oocytes transfected with Cav1.2 and Cavβ2a, and the stoichiometry of Ca2+/CaM: CT was 2:1. Furthermore, they suggested that an interaction between the Ca2+/C-lobe and preIQ C-region is involved in CDF [107] (see also [104]).

5. Channel Clustering/Multimerization

Synchronized openings of multiple Cav1.2 channels have been reported [220]. These findings indicate that CaM antagonists (e.g., W7 and CaM inhibitory peptide), activators of protein kinase Cα, or a specific Cav1.2 channel mutation (associated with Timothy syndrome) that moves CaM away from the IQ domain in the C-terminal tail of the channel promoted coupled gating between Cav1.2 channels [220].
Fallon et al. reported the crystal structure of Cav1.2, which contains an anti-parallel dimer of preIQ helices bridged by two Ca2+/CaMs (see above) [127]. However, sedimentation equilibrium experiments using purified Ca2+/CaM:Cav1.2 preIQ–IQ complexes suggested a model of 2:1 stoichiometry [107]. Furthermore, the measurement of the discrete photobleaching steps of a Cav1.2 a1 subunit bearing a green fluorescent protein tag in its C-terminal revealed that most channels (>95%) bleached in a single step, suggesting that most channels exist in monomeric form [107,221].
In contrast, Dixon et al. demonstrated that Cav1.2 channels form functional couplings through C-terminal-to-C-terminal interactions, mediated by Ca2+/CaM binding to the preIQ domain in native ventricular myocytes and an in vitro expression system [222]. The same group reported that β-adrenergic stimulation promotes surface expression of a sub-sarcolemmal pool of Cav1.2 channels and formation of channel “superclusters”, thereby enhancing channel co-operability [223].

6. Trafficking of Cav1.2 Channels

Trafficking of membrane proteins, such as Cav channels, includes several processes: (1) exportation from the endoplasmic reticulum; (2) transportation to the cell surface; (3) stabilization in the plasma membrane; and (4) internalization, degradation, and recycling to the plasma membrane [224,225,226]. For Cav1.2 channels, these processes are regulated by the intrinsic properties of isoforms of the α1 subunit of the channel, the associated auxiliary subunits, and the regulatory proteins such as CaM.
Ca2+/CaM reportedly binds to GDP-bound inactive RGK proteins and subsequently promotes transformation of RGK to the GTP-bound, Ca2+/CaM-released activated form, which antagonizes binding of Cavβ to the Cavα1 subunit by competitive binding of Cavβ to RGK [227,228,229,230,231]. Thus, the overall effect of Ca2+/CaM is the inhibition of Ca2+ channel cell-surface expression. However, activity induced Ca2+/CaM (but not apoCaM) reportedly accelerated the trafficking of channels to the neuronal cell surface [232]. In addition, overexpression of Ca2+-insensitive CaM1234 accelerated the trafficking of channels to the surface of cultured cells, similar to native CaM [233]. Gao et al. reported that the surface expression of mutant Cav1.2 channels lacking the preIQ3 region (i.e., preIQ-C region) was completely abolished, despite the presence of the β2a subunit, suggesting that the binding of CaM to the channel is required for its surface expression [234], see also [232,235]. Furthermore, reports suggest that α-actinin binds to the IQ domain and stabilizes Cav1.2 at the surface membrane, and its displacement by Ca2+/CaM induces Ca2+-dependent endocytosis of these channels [158,236]. Thus, the role of CaM in trafficking of Cav1.2 channels remains controversial and further studies are required.

7. Conclusions and Perspectives

Many questions about the regulation of Cav1.2 channels by CaM remain to be answered. For example, CaM and other intracellular proteins that interact with channels, such as CaBP1 and α-actinin, may contribute to channel repriming [163]. In the future, it will be interesting to determine whether these proteins can prevent rundown of Cav1.2 channels in inside-out patches. In addition, CDF seems to be mediated by various mechanisms, including one mediated by the direct CaM interaction with channels and other mechanisms not related to the direct interaction of CaM and the channel. Reportedly, CaM can indirectly influence channel activity through activation of CaMKII, phosphodiesterase type 1, or calcineurin, for reviews see [237]. In particular, the CaMKII-mediated phosphorylation of channels enhances their activity, contributing to CDF, therefore, each mechanism should be analyzed independently. For CDI, multiple conformations of the channel-CaM complex in its inactivated state can exist, i.e., bridging of the N- and C-terminals by CaM and a conformation formed by two CaMs at the preIQ-IQ region of the C-terminus. From this viewpoint, solving the structure of the intracellular regions of the Cav1.2 channel would be an important achievement. Finally, although a role for CaM seems plausible for the clustering and trafficking of Cav1.2 channels, more studies are required to reach a clear conclusion.

Author Contributions

Conceptualization and original draft preparation, M.K.; review and editing, E.M., D.S., J.X., Q.G. and L.H. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by grants from the JSPS KAKENHI Grants of Japan (19K16493, 19K07285, 20K07263, 22K06845) and the National Natural Science Foundation of China (31471091, 32271160).

Institutional Review Board Statement

The animal study protocols were approved by the Institutional Review Board of Kagoshima University and China Medical University (1991–2022).

Acknowledgments

The authors are grateful to the members of our laboratories who took part in the studies presented in this article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kakiuchi, S.; Yamazaki, R. Calcium Dependent Phosphodiesterase Activity and Its Activating Factor (PAF) from Brain Studies on Cyclic 3’,5’-Nucleotide Phosphodiesterase (III). Biochem. Biophys. Res. Commun. 1970, 41, 1104–1110. [Google Scholar] [CrossRef] [PubMed]
  2. Halling, D.B.; Liebeskind, B.J.; Hall, A.W.; Aldrich, R.W. Conserved Properties of Individual Ca2+-Binding Sites in Calmodulin. Proc. Natl. Acad. Sci. USA 2016, 113, E1216–E1225. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Hoeflich, K.P.; Ikura, M. Calmodulin in Action: Diversity in Target Recognition and Activation Mechanisms. Cell 2002, 108, 739–742. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Villalobo, A.; Ishida, H.; Vogel, H.J.; Berchtold, M.W. Calmodulin as a Protein Linker and a Regulator of Adaptor/Scaffold Proteins. Biochim. Biophys. Acta Mol. Cell Res. 2018, 1865, 507–521. [Google Scholar] [CrossRef] [PubMed]
  5. Babu, Y.S.; Sack, J.S.; Greenhough, T.J.; Bugg, C.E.; Means, A.R.; Cook, W.J. Three-Dimensional Structure of Calmodulin. Nature 1985, 315, 37–40. [Google Scholar] [CrossRef] [PubMed]
  6. Rhoads, A.R.; Friedberg, F. Sequence motifs for calmodulin recognition. FEBS J. 1997, 11, 331–340. [Google Scholar] [CrossRef] [PubMed]
  7. Catterall, W.A. Voltage-Gated Calcium Channels. Cold Spring Harb. Perspect. Biol. 2011, 3, a003947. [Google Scholar] [CrossRef]
  8. Hofmann, F.; Flockerzi, V.; Kahl, S.; Wegener, J.W. L-Type CaV1.2 Calcium Channels: From in Vitro Findings to in Vivo Function. Physiol. Rev. 2014, 94, 303–326. [Google Scholar] [CrossRef] [Green Version]
  9. Kovalevskaya, N.V.; Van De Waterbeemd, M.; Bokhovchuk, F.M.; Bate, N.; Bindels, R.J.M.; Hoenderop, J.G.J.; Vuister, G.W. Structural Analysis of Calmodulin Binding to Ion Channels Demonstrates the Role of Its Plasticity in Regulation. Pflugers Arch. 2013, 465, 1507–1519. [Google Scholar] [CrossRef]
  10. Sorensen, A.B.; Søndergaard, M.T.; Overgaard, M.T. Calmodulin in a Heartbeat. FEBS J. 2013, 280, 5511–5532. [Google Scholar] [CrossRef]
  11. Ben-Johny, M.; Yue, D.T. Calmodulin Regulation (Calmodulation) of Voltage-Gated Calcium Channels. J. Gen. Physiol. 2014, 143, 679–692. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Striessnig, J.; Pinggera, A.; Kaur, G.; Bock, G.; Tuluc, P. L-Type Ca2+ Channels in Heart and Brain. Wiley Interdiscip. Rev. Membr. Transp. Signal 2014, 3, 15–38. [Google Scholar] [CrossRef] [PubMed]
  13. Simms, B.A.; Zamponi, G.W. Neuronal Voltage-Gated Calcium Channels: Structure, Function, and Dysfunction. Neuron 2014, 82, 24–45. [Google Scholar] [CrossRef] [Green Version]
  14. Dolphin, A.C. Voltage-Gated Calcium Channels and Their Auxiliary Subunits: Physiology and Pathophysiology and Pharmacology. J. Physiol. 2016, 594, 5369–5390. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Ames, J.B. L-Type Ca2+ Channel Regulation by Calmodulin and CaBP1. Biomolecules 2021, 11, 1811. [Google Scholar] [CrossRef] [PubMed]
  16. Hagiwara, S.; Byerly, L. The Calcium Channel. Trends Neurosci. 1983, 6, 189–193. [Google Scholar] [CrossRef]
  17. Kostyuk, P.G. Intracellular Perfusion of Nerve Cells and Its Effects on Membrane Currents. Physiol. Rev. 1984, 64, 435–454. [Google Scholar] [CrossRef]
  18. Mcdonald, T.F.; Pelzer, S.; Trautwein, W.; Pelzer, D.J. Regulation and Modulation of Calcium Channels in Cardiac, Skeletal, and Smooth Muscle Cells. Physiol. Rev. 1994, 74, 365–507. [Google Scholar] [CrossRef]
  19. Fedulova, S.A.; Kostyuk, P.G.; Veselovsky, N.S. Calcium Channels in the Somatic Membrane of the Rat Dorsal Root Ganglion Neurons, Effect of CAMP. Brain Res. 1981, 214, 210–214. [Google Scholar] [CrossRef]
  20. Chad, J.E.; Eckert, R. An Enzymatic Mechanism for Calcium Current Inactivation in Dialysed Helix Neurones. J. Physiol. 1986, 378, 31–51. [Google Scholar] [CrossRef]
  21. Armstrong, D.; Eckert, R. Voltage-Activated Calcium Channels That Must Be Phosphorylated to Respond to Membrane Depolarization. Proc. Natl. Acad. Sci. USA 1987, 84, 2518–2522. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Ono, K.; Fozzard, H.A. Phosphorylation Restores Activity of L-Type Calcium Channels after Rundown in inside-out Patches from Rabbit Cardiac Cells. J. Physiol. 1992, 454, 673–688. [Google Scholar] [CrossRef] [PubMed]
  23. Irisawa, H.; Kokubun, S. Modulation by Intracellular ATP and Cyclic AMP of the Slow Inward Current in Isolated Single Ventricular Cells of the Guinea-Pig. J. Physiol. 1983, 338, 321–337. [Google Scholar] [CrossRef]
  24. Kameyama, M.; Kameyama, A.; Takano, E.; Maki, M. Run-down of the Cardiac L-Type Ca2+ Channel: Partial Restoration of Channel Activity in Cell-Free Patches by Calpastatin. Pflugers Arch. 1998, 435, 344–349. [Google Scholar] [CrossRef] [PubMed]
  25. Kameyama, M.; Kameyama, A.; Nakayama, T.; Kaibara, M. Tissue Extract Recovers Cardiac Calcium Channels from “Run-Down”. Pflugers Arch. 1988, 412, 328–330. [Google Scholar] [CrossRef] [PubMed]
  26. Kameyama, A.; Yazawa, K.; Kaibara, M.; Ozono, K.; Kameyama, M. Run-down of the Cardiac Ca2+ Channel: Characterization and Restoration of Channel Activity by Cytoplasmic Factors. Pflugers Arch. 1997, 433, 547–556. [Google Scholar] [CrossRef] [PubMed]
  27. Romanin, C.; Grösswagen, P.; Schindler, H. Calpastatin and Nucleotides Stabilize Cardiac Calcium Channel Activity in Excised Patches. Pflugers Arch. 1991, 418, 86–92. [Google Scholar] [CrossRef] [PubMed]
  28. Hao, L.Y.; Kameyama, A.; Kameyama, M. A Cytoplasmic Factor, Calpastatin and ATP Together Reverse Run-down of Ca2+ Channel Activity in Guinea-Pig Heart. J. Physiol. 1999, 514 Pt 3, 687–699. [Google Scholar] [CrossRef]
  29. Belles, B.; Hescheler, J.; Trautwein, W.; Blomgren, K.; Karlsson, J.O. A Possible Physiological Role of the Ca-Dependent Protease Calpain and Its Inhibitor Calpastatin on the Ca Current in Guinea Pig Myocytes. Pflugers Arch. 1988, 412, 554–556. [Google Scholar] [CrossRef]
  30. Horn, R.; Marty, A. Muscarinic Activation of Ionic Currents Measured by a New Whole-Cell Recording Method. J. Gen Physiol. 1988, 92, 145–159. [Google Scholar] [CrossRef]
  31. Hao, L.Y.; Kameyama, A.; Kuroki, S.; Takano, J.; Takano, E.; Maki, M.; Kameyama, M. Calpastatin Domain L Is Involved in the Regulation of L-Type Ca2+ Channels in Guinea Pig Cardiac Myocytes. Biochem. Biophys. Res. Commun. 2000, 279, 756–761. [Google Scholar] [CrossRef]
  32. Minobe, E.; Hao, L.Y.; Saud, Z.A.; Xu, J.J.; Kameyama, A.; Maki, M.; Jewell, K.K.; Parr, T.; Bardsley, R.G.; Kameyama, M. A Region of Calpastatin Domain L That Reprimes Cardiac L-Type Ca2+ Channels. Biochem. Biophys. Res. Commun. 2006, 348, 288–294. [Google Scholar] [CrossRef] [PubMed]
  33. Xu, J.J.; Hao, L.Y.; Kameyama, A.; Kameyama, M. Calmodulin Reverses Rundown of L-Type Ca2+ Channels in Guinea Pig Ventricular Myocytes. Am. J. Physiol. Cell Physiol. 2004, 287, C1717–C1724. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Saud, Z.A.; Minobe, E.; Wang, W.Y.; Han, D.Y.; Horiuchi, M.; Hao, L.Y.; Kameyama, M. Calpastatin Binds to a Calmodulin-Binding Site of Cardiac Cav1.2 Ca2+ Channels. Biochem. Biophys. Res. Commun. 2007, 364, 372–377. [Google Scholar] [CrossRef] [PubMed]
  35. Minobe, E.; Asmara, H.; Saud, Z.A.; Kameyama, M. Calpastatin Domain L Is a Partial Agonist of the Calmodulin-Binding Site for Channel Activation in Cav1.2 Ca2+ Channels. J. Biol. Chem. 2011, 286, 39013–39022. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Lee, A.; Wong, S.T.; Gallagher, D.; Li, B.; Storm, D.R.; Scheuer, T.; Catterall, W.A. Ca2+/Calmodulin Binds to and Modulates P/Q-Type Calcium Channels. Nature 1999, 399, 155–159. [Google Scholar] [CrossRef] [PubMed]
  37. Lee, A.; Scheuer, T.; Catterall, W.A. Ca2+/Calmodulin-Dependent Facilitation and Inactivation of P/Q-Type Ca2+ Channels. J. Neurosci. 2000, 20, 6830–6838. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Mouton, J.; Feltz, A.; Maulet, Y. Interactions of Calmodulin with Two Peptides Derived from the C-Terminal Cytoplasmic Domain of the Cav1.2 Ca2+ Channel Provide Evidence for a Molecular Switch Involved in Ca2+-Induced Inactivation. J. Biol. Chem. 2001, 276, 22359–22367. [Google Scholar] [CrossRef] [Green Version]
  39. Pitt, G.S.; Zühlke, R.D.; Hudmon, A.; Schulman, H.; Reuter, H.; Tsien, R.W. Molecular Basis of Calmodulin Tethering and Ca2+-Dependent Inactivation of L-Type Ca2+ Channels. J. Biol. Chem. 2001, 276, 30794–30802. [Google Scholar] [CrossRef] [Green Version]
  40. DeMaria, C.D.; Soong, T.W.; Alseikhan, B.A.; Alvania, R.S.; Yue, D.T. Calmodulin Bifurcates the Local Ca2+ Signal That Modulates P/Q-Type Ca2+ Channels. Nature 2001, 411, 484–489. [Google Scholar] [CrossRef]
  41. Zühlke, R.D.; Pitt, G.S.; Deisseroth, K.; Tsien, R.W.; Reuter, H. Calmodulin Supports Both Inactivation and Facilitation of L-Type Calcium Channels. Nature 1999, 399, 159–162. [Google Scholar] [CrossRef] [PubMed]
  42. Zühlke, R.D.; Pitt, G.S.; Tsien, R.W.; Reuter, H. Ca2+ -Sensitive Inactivation and Facilitation of L-Type Ca2+ Channels Both Depend on Specific Amino Acid Residues in a Consensus Calmodulin-Binding Motif in the α1C Subunit. J. Biol. Chem. 2000, 275, 21121–21129. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Evans, T.I.A.; Hell, J.W.; Shea, M.A. Thermodynamic Linkage between Calmodulin Domains Binding Calcium and Contiguous Sites in the C-Terminal Tail of CaV1.2. Biophys. Chem. 2011, 159, 172–187. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Turner, M.; Anderson, D.E.; Bartels, P.; Nieves-Cintron, M.; Coleman, A.M.; Henderson, P.B.; Man, K.N.M.; Tseng, P.; Yarovoy, V.; Bers, D.M.; et al. α-Actinin-1 Promotes Activity of the L-Type Ca2+ Channel Cav 1.2. EMBO J. 2020, 39, e102622. [Google Scholar] [CrossRef]
  45. Wu, X.; Bers, D.M. Free and Bound Intracellular Calmodulin Measurements in Cardiac Myocytes. Cell Calcium 2007, 41, 353–364. [Google Scholar] [CrossRef] [Green Version]
  46. Shao, D.; Zhao, M.; Xu, J.; Feng, R.; Guo, F.; Hu, H.; Sun, X.; Gao, Q.; He, G.; Sun, W.; et al. The Individual N- and C-Lobes of Calmodulin Tether to the Cav1.2 Channel and Rescue the Channel Activity from Run-down in Ventricular Myocytes of Guinea-Pig Heart. FEBS Lett. 2014, 588, 3855–3861. [Google Scholar] [CrossRef] [Green Version]
  47. Bartels, P.; Salveson, I.; Coleman, A.M.; Anderson, D.E.; Jeng, G.; Estrada-Tobar, Z.M.; Man, K.N.M.; Yu, Q.; Kuzmenkina, E.; Nieves-Cintron, M.; et al. Half-Calcified Calmodulin Promotes Basal Activity and Inactivation of the L-Type Calcium Channel CaV1.2. J. Biol. Chem. 2022, 298, 102701. [Google Scholar] [CrossRef]
  48. Adams, P.J.; Ben-Johny, M.; Dick, I.E.; Inoue, T.; Yue, D.T. Apocalmodulin Itself Promotes Ion Channel Opening and Ca2+ Regulation. Cell 2014, 159, 608–622. [Google Scholar] [CrossRef] [Green Version]
  49. Banerjee, R.; Yoder, J.B.; Yue, D.T.; Amzel, L.M.; Tomaselli, G.F.; Gabelli, S.B.; Ben-Johny, M. Bilobal Architecture Is a Requirement for Calmodulin Signaling to CaV1.3 Channels. Proc. Natl. Acad. Sci. USA 2018, 115, E3026–E3035. [Google Scholar] [CrossRef] [Green Version]
  50. Liu, X.; Yang, P.S.; Yang, W.; Yue, D.T. Enzyme-Inhibitor-like Tuning of Ca2+ Channel Connectivity with Calmodulin. Nature 2010, 463, 968–972. [Google Scholar] [CrossRef] [Green Version]
  51. Han, D.Y.; Minobe, E.; Wang, W.Y.; Guo, F.; Xu, J.J.; Hao, L.Y.; Kameyama, M. Calmodulin- and Ca2+-Dependent Facilitation and Inactivation of the Cav1.2 Ca2+ Channels in Guinea-Pig Ventricular Myocytes. J. Pharmacol. Sci. 2010, 112, 310–319. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Kobayashi, T.; Yamada, Y.; Fukao, M.; Shiratori, K.; Tsutsuura, M.; Tanimoto, K.; Tohse, N. The GK Domain of the Voltage-Dependent Calcium Channel β Subunit Is Essential for Binding to the α Subunit. Biochem. Biophys. Res. Commun. 2007, 360, 679–683. [Google Scholar] [CrossRef] [PubMed]
  53. Liu, G.; Papa, A.; Katchman, A.N.; Zakharov, S.I.; Roybal, D.; Hennessey, J.A.; Kushner, J.; Yang, L.; Chen, B.X.; Kushnir, A.; et al. Mechanism of Adrenergic CaV1.2 Stimulation Revealed by Proximity Proteomics. Nature 2020, 577, 695–700. [Google Scholar] [CrossRef] [PubMed]
  54. Yazawa, K.; Kameyama, A.; Yasui, K.; Li, J.M.; Kameyama, M. ATP Regulates Cardiac Ca2+ Channel Activity via a Mechanism Independent of Protein Phosphorylation. Pflugers Arch. 1997, 433, 557–562. [Google Scholar] [CrossRef]
  55. Liu, S.Y.; Xu, J.J.; Minobe, E.; Gao, Q.H.; Feng, R.; Zhao, M.M.; Guo, F.; Yang, L.; Hao, L.Y.; Kameyama, M. Nucleotides Maintain the Activity of Cav1.2 Channels in Guinea-Pig Ventricular Myocytes. Biochem. Biophys. Res. Commun. 2015, 460, 813–818. [Google Scholar] [CrossRef]
  56. O’Rourke, B.; Backx, P.H.; Marban, E. Phosphorylation-Independent Modulation of L-Type Calcium Channels by Magnesium-Nucleotide Complexes. Science 1992, 257, 245–248. [Google Scholar] [CrossRef]
  57. Feng, R.; Xu, J.; Minobe, E.; Kameyama, A.; Yang, L.; Yu, L.; Hao, L.; Kameyama, M. Adenosine Triphosphate Regulates the Activity of Guinea Pig Cav1.2 Channel by Direct Binding to the Channel in a Dose-Dependent Manner. Am. J. Physiol. Cell Physiol. 2014, 306, C856–C863. [Google Scholar] [CrossRef] [Green Version]
  58. Hao, L.Y.; Xu, J.J.; Minobe, E.; Kameyama, A.; Kameyama, M. Calmodulin Kinase II Activation Is Required for the Maintenance of Basal Activity of L-Type Ca2+ Channels in Guinea-Pig Ventricular Myocytes. J. Pharmacol. Sci. 2008, 108, 290–300. [Google Scholar] [CrossRef] [Green Version]
  59. Hao, L.Y.; Wang, W.Y.; Minobe, E.; Han, D.Y.; Xu, J.J.; Kameyama, A.; Kameyama, M. The Distinct Roles of Calmodulin and Calmodulin Kinase II in the Reversal of Run-down of L-Type Ca2+ Channels in Guinea-Pig Ventricular Myocytes. J. Pharmacol. Sci. 2009, 111, 416–425. [Google Scholar] [CrossRef] [Green Version]
  60. Wang, W.Y.; Hao, L.Y.; Minobe, E.; Saud, Z.A.; Han, D.Y.; Kameyama, M. CaMKII Phosphorylates a Threonine Residue in the C-Terminal Tail of Cav1.2 Ca2+ Channel and Modulates the Interaction of the Channel with Calmodulin. J. Physiol. Sci. 2009, 59, 283–290. [Google Scholar] [CrossRef]
  61. Lei, M.; Xu, J.; Gao, Q.; Minobe, E.; Kameyama, M.; Hao, L. PKA Phosphorylation of Cav1.2 Channel Modulates the Interaction of Calmodulin with the C Terminal Tail of the Channel. J. Pharmacol. Sci. 2018, 137, 187–194. [Google Scholar] [CrossRef] [PubMed]
  62. Halling, D.B.; Aracena-Parks, P.; Hamilton, S.L. Regulation of Voltage-Gated Ca2+ Channels by Calmodulin. Sci. STKE 2006, 2006, er1. [Google Scholar] [CrossRef]
  63. Chapman, R.A.; Niedergerke, R. Interaction between Heart Rate and Calcium Concentration in the Control of Contractile Strength of the Frog Heart. J. Physiol. 1970, 211, 423–443. [Google Scholar] [CrossRef]
  64. Balcazar, D.; Regge, V.; Santalla, M.; Meyer, H.; Paululat, A.; Mattiazzi, A.; Ferrero, P. SERCA Is Critical to Control the Bowditch Effect in the Heart. Sci. Rep. 2018, 8, 12447. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Noble, S.; Shimoni, Y. The Calcium and Frequency Dependence of the Slow Inward Current ‘Staircase’ in Frog Atrium. J. Physiol. 1981, 310, 57–75. [Google Scholar] [CrossRef] [PubMed]
  66. Noble, S.; Shimoni, Y. Voltage-Dependent Potentiation of the Slow Inward Current in Frog Atrium. J. Physiol. 1981, 310, 77–95. [Google Scholar] [CrossRef] [Green Version]
  67. Marban, E.; Tsien, R.W. Enhancement of Calcium Current during Digitalis Inotropy in Mammalian Heart: Positive Feed-Back Regulation by Intracellular Calcium? J. Physiol. 1982, 329, 589–614. [Google Scholar] [CrossRef]
  68. Lee, K.S. Potentiation of the Calcium-Channel Currents of Internally Perfused Mammalian Heart Cells by Repetitive Depolarization. Proc. Natl. Acad. Sci. USA 1987, 84, 3941–3945. [Google Scholar] [CrossRef] [Green Version]
  69. Fedida, D.; Noble, D.; Spindler, A.J. Mechanism of the Use Dependence of Ca2+ Current in Guinea-Pig Myocytes. J. Physiol. 1988, 405, 461–475. [Google Scholar] [CrossRef] [Green Version]
  70. Tseng, G.N. Calcium Current Restitution in Mammalian Ventricular Myocytes Is Modulated by Intracellular Calcium. Circ. Res. 1988, 63, 468–482. [Google Scholar] [CrossRef] [Green Version]
  71. Zygmunt, A.C.; Maylie, J. Stimulation-Dependent Facilitation of the High Threshold Calcium Current in Guinea-Pig Ventricular Myocytes. J. Physiol. 1990, 428, 653–671. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Pietrobon, D.; Hess, P. Novel Mechanism of Voltage-Dependent Gating in L-Type Calcium Channels. Nature 1990, 346, 651–655. [Google Scholar] [CrossRef]
  73. Kleppisch, T.; Pedersen, K.; Strubing, C.; Bosse-Doenecke, E.; Flockerzi, V.; Hofmann, F.; Hescheler, J. Double-Pulse Facilitation of Smooth Muscle α1-Subunit Ca2+ Channels Expressed in CHO Cells. EMBO J. 1994, 13, 2502–2507. [Google Scholar] [CrossRef] [PubMed]
  74. Gurney, A.M.; Charnet, P.; Pye, J.M.; Nargeot, J. Augmentation of Cardiac Calcium Current by Flash Photolysis of Intracellular Caged-Ca2+ Molecules. Nature 1989, 341, 65–68. [Google Scholar] [CrossRef]
  75. Hadley, R.W.; Lederer, W.J. Ca2+ and Voltage Inactivate Ca2+ Channels in Guinea-Pig Ventricular Myocytes through Independent Mechanisms. J. Physiol. 1991, 444, 257–268. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Bates, S.E.; Gurney, A.M. Ca2+-Dependent Block and Potentiation of L-Type Calcium Current in Guinea-Pig Ventricular Myocytes. J. Physiol. 1993, 466, 345–365. [Google Scholar] [CrossRef]
  77. Hirano, Y.; Hiraoka, M. Dual Modulation of Unitary L-Type Ca2+ Channel Currents by [Ca2+]i in Fura-2-Loaded Guinea-Pig Ventricular Myocytes. J. Physiol. 1994, 480, 449–463. [Google Scholar] [CrossRef]
  78. Yamaoka, K.; Seyama, I. Regulation of Ca Channel by Intracellular Ca2+ and Mg2+ in Frog Ventricular Cells. Pflugers Arch. 1996, 431, 305–317. [Google Scholar] [CrossRef]
  79. Nie, H.G.; Hao, L.Y.; Xu, J.J.; Minobe, E.; Kameyama, A.; Kameyama, M. Distinct Roles of CaM and Ca2+/CaM-Dependent Protein Kinase II in Ca2+-Dependent Facilitation and Inactivation of Cardiac L-Type Ca2+ Channels. J. Physiol. Sci. 2008, 046. [Google Scholar] [CrossRef] [Green Version]
  80. Anderson, M.E.; Braun, A.P.; Schulman, H.; Premack, B.A. Multifunctional Ca2+/Calmodulin-Dependent Protein Kinase Mediates Ca2+-Induced Enhancement of the L-Type Ca2+ Current in Rabbit Ventricular Myocytes. Circ. Res. 1994, 75, 854–861. [Google Scholar] [CrossRef] [Green Version]
  81. Xiao, R.P.; Cheng, H.; Lederer, W.J.; Suzuki, T.; Lakatta, E.G. Dual Regulation of Ca2+/Calmodulin-Dependent Kinase II Activity by Membrane Voltage and by Calcium Influx. Proc. Natl. Acad. Sci. USA 1994, 91, 9659–9663. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Yuan, W.; Bers, D.M. Ca-Dependent Facilitation of Cardiac Ca Current Is Due to Ca-Calmodulin-Dependent Protein Kinase. Am. J. Physiol. 1994, 267, H982–H993. [Google Scholar] [CrossRef] [PubMed]
  83. Hudmon, A.; Schulman, H.; Kim, J.; Maltez, J.M.; Tsien, R.W.; Pitt, G.S. CaMKII Tethers to L-Type Ca2+ Channels, Establishing a Local and Dedicated Integrator of Ca2+ Signals for Facilitation. J. Cell Biol. 2005, 171, 537–547. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Abiria, S.A.; Colbran, R.J. CaMKII Associates with CaV1.2 L-Type Calcium Channels via Selected β Subunits to Enhance Regulatory Phosphorylation. J. Neurochem. 2010, 112, 150–161. [Google Scholar] [CrossRef] [Green Version]
  85. Erxleben, C.; Liao, Y.; Gentile, S.; Chin, D.; Gomez-Alegria, C.; Mori, Y.; Birnbaumer, L.; Armstrong, D.L. Cyclosporin and Timothy Syndrome Increase Mode 2 Gating of CaV1.2 Calcium Channels through Aberrant Phosphorylation of S6 Helices. Proc. Natl. Acad. Sci. USA 2006, 103, 3932–3937. [Google Scholar] [CrossRef] [Green Version]
  86. Lee, T.S.; Karl, R.; Moosmang, S.; Lenhardt, P.; Klugbauer, N.; Hofmann, F.; Kleppisch, T.; Welling, A. Calmodulin Kinase II Is Involved in Voltage-Dependent Facilitation of the L-Type Cav1.2 Calcium Channel: Identification of the Phosphorylation Sites. J. Biol. Chem. 2006, 281, 25560–25567. [Google Scholar] [CrossRef] [Green Version]
  87. Grueter, C.E.; Abiria, S.A.; Dzhura, I.; Wu, Y.; Ham, A.J.L.; Mohler, P.J.; Anderson, M.E.; Colbran, R.J. L-Type Ca2+ Channel Facilitation Mediated by Phosphorylation of the Beta Subunit by CaMKII. Mol. Cell 2006, 23, 641–650. [Google Scholar] [CrossRef]
  88. Blaich, A.; Welling, A.; Fischer, S.; Wegener, J.W.; Köstner, K.; Hofmann, F.; Moosmang, S. Facilitation of Murine Cardiac L-Type Cav1.2 Channel Is Modulated by Calmodulin Kinase II-Dependent Phosphorylation of S1512 and S1570. Proc. Natl. Acad. Sci. USA 2010, 107, 10285–10289. [Google Scholar] [CrossRef] [Green Version]
  89. Brandmayr, J.; Poomvanicha, M.; Domes, K.; Ding, J.; Blaich, A.; Wegener, J.W.; Moosmang, S.; Hofmann, F. Deletion of the C-Terminal Phosphorylation Sites in the Cardiac β-Subunit Does Not Affect the Basic β-Adrenergic Response of the Heart and the Cav1.2 Channel. J. Biol. Chem. 2012, 287, 22584–22592. [Google Scholar] [CrossRef] [Green Version]
  90. Wang, S.; Li, J.; Liu, Y.; Zhang, J.; Zheng, X.; Sun, X.; Lei, S.; Kang, Z.; Chen, X.; Lei, M.; et al. Distinct Roles of Calmodulin and Ca2+/Calmodulin-Dependent Protein Kinase II in Isopreterenol-Induced Cardiac Hypertrophy. Biochem. Biophys. Res. Commun. 2020, 526, 960–966. [Google Scholar] [CrossRef]
  91. Zheng, X.; Su, F.; Kang, Z.; Li, J.; Zhang, C.; Zhang, Y.; Hao, L. Analysis of Therapeutic Targets of A Novel Peptide Athycaltide-1 in the Treatment of Isoproterenol-Induced Pathological Myocardial Hypertrophy. Cardiovasc. Ther. 2022, 2022, 2715084. [Google Scholar] [CrossRef]
  92. Van Petegem, F.; Chatelain, F.C.; Minor, D.L. Insights into Voltage-Gated Calcium Channel Regulation from the Structure of the CaV1.2 IQ Domain-Ca2+/Calmodulin Complex. Nat. Struct. Mol. Biol. 2005, 12, 1108–1115. [Google Scholar] [CrossRef] [PubMed]
  93. Yang, D.; Song, L.S.; Zhu, W.Z.; Chakir, K.; Wang, W.; Wu, C.; Wang, Y.; Xiao, R.P.; Chen, S.R.W.; Cheng, H. Calmodulin Regulation of Excitation-Contraction Coupling in Cardiac Myocytes. Circ. Res. 2003, 92, 659–667. [Google Scholar] [CrossRef] [Green Version]
  94. Wu, Y.; Dzhura, I.; Colbran, R.J.; Anderson, M.E. Calmodulin Kinase and a Calmodulin-Binding ‘IQ’ Domain Facilitate L-Type Ca2+ Current in Rabbit Ventricular Myocytes by a Common Mechanism. J. Physiol. 2001, 535, 679–687. [Google Scholar] [CrossRef]
  95. Rodney, G.G.; Williams, B.Y.; Strasburg, G.M.; Beckingham, K.; Hamilton, S.L. Regulation of RYR1 Activity by Ca2+ and Calmodulin. Biochemistry 2000, 39, 7807–7812. [Google Scholar] [CrossRef]
  96. Guo, F.; Minobe, E.; Yazawa, K.; Asmara, H.; Bai, X.Y.; Han, D.Y.; Hao, L.Y.; Kameyama, M. Both N- and C-Lobes of Calmodulin Are Required for Ca2+-Dependent Regulations of CaV1.2 Ca2+ Channels. Biochem. Biophys. Res. Commun. 2010, 391, 1170–1176. [Google Scholar] [CrossRef] [PubMed]
  97. Kim, J.; Ghosh, S.; Nunziato, D.A.; Pitt, G.S. Identification of the Components Controlling Inactivation of Voltage-Gated Ca2+ Channels. Neuron 2004, 41, 745–754. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Dick, I.E.; Tadross, M.R.; Liang, H.; Tay, L.H.; Yang, W.; Yue, D.T. A Modular Switch for Spatial Ca2+ Selectivity in the Calmodulin Regulation of CaV Channels. Nature 2008, 451, 830–834. [Google Scholar] [CrossRef] [PubMed]
  99. Yadav, D.K.; Anderson, D.E.; Hell, J.W.; Ames, J.B. Calmodulin Promotes a Ca2+-Dependent Conformational Change in the C-Terminal Regulatory Domain of CaV 1.2. FEBS Lett. 2022, 596, 2974–2985. [Google Scholar] [CrossRef]
  100. Maier, L.S.; Ziolo, M.T.; Bossuyt, J.; Persechini, A.; Mestril, R.; Bers, D.M. Dynamic Changes in Free Ca-Calmodulin Levels in Adult Cardiac Myocytes. J. Mol. Cell Cardiol. 2006, 41, 451–458. [Google Scholar] [CrossRef]
  101. Zhao, M.; Feng, R.; Shao, D.; Liu, S.; Lei, M.; Wang, H.; Sun, X.; Guo, F.; Hu, H.; Kameyama, M.; et al. Mg2+-Dependent Facilitation and Inactivation of L-Type Ca2+ Channels in Guinea Pig Ventricular Myocytes. J. Pharmacol. Sci. 2015, 129, 143–149. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Peterson, B.Z.; DeMaria, C.D.; Yue, D.T. Calmodulin Is the Ca2+ Sensor for Ca2+ -Dependent Inactivation of L-Type Calcium Channels. Neuron 1999, 22, 549–558. [Google Scholar] [CrossRef] [Green Version]
  103. Zühlke, R.D.; Reuter, H. Ca2+-Sensitive Inactivation of L-Type Ca2+ Channels Depends on Multiple Cytoplasmic Amino Acid Sequences of the Alpha1C Subunit. Proc. Natl. Acad. Sci. USA 1998, 95, 3287–3294. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Young Kim, E.; Findeisen, F.; Minor, D.L. Calmodulin Interactions with Cav1 and Cav2 Voltage-Gated Calcium Channel IQ Domains. In Handbook of Metalloproteins; John Wiley: Hoboken, NJ, USA, 2010. [Google Scholar] [CrossRef] [Green Version]
  105. Ben-Johny, M.; Dick, I.E.; Sang, L.; Limpitikul, W.B.; Kang, P.W.; Niu, J.; Banerjee, R.; Yang, W.; Babich, J.S.; Issa, J.B.; et al. Towards a Unified Theory of Calmodulin Regulation (Calmodulation) of Voltage-Gated Calcium and Sodium Channels. Curr. Mol. Pharmacol. 2015, 8, 188–205. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Simms, B.A.; Souza, I.A.; Zamponi, G.W. A Novel Calmodulin Site in the Cav1.2 N-Terminus Regulates Calcium-Dependent Inactivation. Pflugers Arch. 2014, 466, 1793–1803. [Google Scholar] [CrossRef] [PubMed]
  107. Kim, E.Y.; Rumpf, C.H.; Van Petegem, F.; Arant, R.J.; Findeisen, F.; Cooley, E.S.; Isacoff, E.Y.; Minor, D.L. Multiple C-Terminal Tail Ca2+/CaMs Regulate Cav1.2 Function but Do Not Mediate Channel Dimerization. EMBO J. 2010, 29, 3924–3938. [Google Scholar] [CrossRef] [Green Version]
  108. Ross, J.; Miura, T.; Kambayashi, M.; Eising, G.P.; Ryu, K.H. Adrenergic Control of the Force-Frequency Relation. Circulation 1995, 92, 2327–2332. [Google Scholar] [CrossRef]
  109. Brehm, P.; Eckert, R. Calcium Entry Leads to Inactivation of Calcium Channel in Paramecium. Science 1978, 202, 1203–1206. [Google Scholar] [CrossRef]
  110. Brehm, P.; Eckert, R.; Tillotson, D. Calcium-mediated Inactivation of Calcium Current in Paramecium. J. Physiol. 1980, 306, 193–203. [Google Scholar] [CrossRef]
  111. Tillotson, D. Inactivation of Ca Conductance Dependent on Entry of Ca Ions in Molluscan Neurons. Proc. Natl. Acad. Sci. USA 1979, 76, 1497–1500. [Google Scholar] [CrossRef] [Green Version]
  112. Eckert, R.; Chad, J.E. Inactivation of Ca Channels. Prog. Biophys. Mol. Biol. 1984, 44, 215–267. [Google Scholar] [CrossRef] [PubMed]
  113. Taiakina, V.; Boone, A.N.; Fux, J.; Senatore, A.; Weber-Adrian, D.; Guillemette, J.G.; Spafford, J.D. The Calmodulin-Binding, Short Linear Motif, NSCaTE Is Conserved in L-Type Channel Ancestors of Vertebrate Cav1.2 and Cav1.3 Channels. PLoS ONE 2013, 8, e61765. [Google Scholar] [CrossRef] [Green Version]
  114. Christel, C.; Lee, A. Ca2+-Dependent Modulation of Voltage-Gated Ca2+ Channels. Biochim. Biophys. Acta 2012, 1820, 1243–1252. [Google Scholar] [CrossRef] [Green Version]
  115. Sang, L.; Vieira, D.C.O.; Yue, D.T.; Ben-Johny, M.; Dick, I.E. The Molecular Basis of the Inhibition of CaV1 Calcium-Dependent Inactivation by the Distal Carboxy Tail. J. Biol. Chem. 2021, 296, 100502. [Google Scholar] [CrossRef]
  116. Zhou, J.; Olcese, R.; Qin, N.; Noceti, F.; Birnbaumer, L.; Stefani, E. Feedback Inhibition of Ca2+ Channels by Ca2+ Depends on a Short Sequence of the C Terminus That Does Not Include the Ca2+ -Binding Function of a Motif with Similarity to Ca2+ -Binding Domains. Proc. Natl. Acad. Sci. USA 1997, 94, 2301–2305. [Google Scholar] [CrossRef] [Green Version]
  117. Qin, N.; Olcese, R.; Bransby, M.; Lin, T.; Birnbaumer, L. Ca2+-Induced Inhibition of the Cardiac Ca2+ Channel Depends on Calmodulin. Proc. Natl. Acad. Sci. USA 1999, 96, 2435–2438. [Google Scholar] [CrossRef] [Green Version]
  118. Soldatov, N.M.; Zühlke, R.D.; Bouron, A.; Reuter, H. Molecular Structures Involved in L-Type Calcium Channel Inactivation. Role of the Carboxyl-Terminal Region Encoded by Exons 40-42 in α1C Subunit in the Kinetics and Ca2+ Dependence of Inactivation. J. Biol. Chem. 1997, 272, 3560–3566. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  119. Oliveria, S.F.; Dittmer, P.J.; Youn, D.H.; Dell’Acqua, M.L.; Sather, W.A. Localized Calcineurin Confers Ca2+-Dependent Inactivation on Neuronal L-Type Ca2+ Channels. J. Neurosci. 2012, 32, 15328–15337. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  120. Oliveria, S.F.; Dell’Acqua, M.L.; Sather, W.A. AKAP79/150 Anchoring of Calcineurin Controls Neuronal L-Type Ca2+ Channel Activity and Nuclear Signaling. Neuron 2007, 55, 261–275. [Google Scholar] [CrossRef] [Green Version]
  121. Hulme, J.T.; Westenforoek, R.E.; Scheuer, T.; Catterall, W.A. Phosphorylation of Serine 1928 in the Distal C-Terminal Domain of Cardiac CaV1.2 Channels during Beta1-Adrenergic Regulation. Proc. Natl. Acad. Sci. USA 2006, 103, 16574–16579. [Google Scholar] [CrossRef] [Green Version]
  122. Dittmer, P.J.; Dell’Acqua, M.L.; Sather, W.A. Ca2+/Calcineurin-Dependent Inactivation of Neuronal L-Type Ca2+ Channels Requires Priming by AKAP-Anchored Protein Kinase A. Cell Rep. 2014, 7, 1410–1416. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. De Leon, M.; Wang, Y.; Jones, L.; Perez-Reyes, E.; Wei, X.; Soong, T.W.; Snutch, T.P.; Yue, D.T. Essential Ca2+-Binding Motif for Ca2+-Sensitive Inactivation of L-Type Ca2+ Channels. Science 1995, 270, 1502–1506. [Google Scholar] [CrossRef] [PubMed]
  124. Soldatov, N.M.; Oz, M.; O’Brien, K.A.; Abernethy, D.R.; Morad, M. Molecular Determinants of L-Type Ca2+ Channel Inactivation. Segment Exchange Analysis of the Carboxyl-Terminal Cytoplasmic Motif Encoded by Exons 40-42 of the Human α1C Subunit Gene. J. Biol. Chem. 1998, 273, 957–963. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Peterson, B.Z.; Lee, J.S.; Mulle, J.G.; Wang, V.; De Leon, M.; Yue, D.T. Critical Determinants of Ca2+-Dependent Inactivation within an EF-Hand Motif of L-Type Ca2+ Channels. Biophys. J. 2000, 78, 1906–1920. [Google Scholar] [CrossRef] [Green Version]
  126. Zhou, H.; Kim, S.A.; Kirk, E.A.; Tippens, A.L.; Sun, H.; Haeseleer, F.; Lee, A. Ca2+-Binding Protein-1 Facilitates and Forms a Postsynaptic Complex with Cav1.2 (L-Type) Ca2+ Channels. J. Neurosci. 2004, 24, 4698–4708. [Google Scholar] [CrossRef] [Green Version]
  127. Fallon, J.L.; Baker, M.R.; Xiong, L.; Loy, R.E.; Yang, G.; Dirksen, R.T.; Hamilton, S.L.; Quiocho, F.A. Crystal Structure of Dimeric Cardiac L-Type Calcium Channel Regulatory Domains Bridged by Ca2+·Calmodulins. Proc. Natl. Acad. Sci. USA 2009, 106, 5135–5140. [Google Scholar] [CrossRef] [Green Version]
  128. Erickson, M.G.; Liang, H.; Mori, M.X.; Yue, D.T. FRET Two-Hybrid Mapping Reveals Function and Location of L-Type Ca2+ Channel CaM Preassociation. Neuron 2003, 39, 97–107. [Google Scholar] [CrossRef] [Green Version]
  129. Poomvanicha, M.; Wegener, J.W.; Blaich, A.; Fischer, S.; Domes, K.; Moosmang, S.; Hofmann, F. Facilitation and Ca2+-Dependent Inactivation Are Modified by Mutation of the Cav1.2 Channel IQ Motif. J. Biol. Chem. 2011, 286, 26702–26707. [Google Scholar] [CrossRef] [Green Version]
  130. Blaich, A.; Pahlavan, S.; Tian, Q.; Oberhofer, M.; Poomvanicha, M.; Lenhardt, P.; Domes, K.; Wegener, J.W.; Moosmang, S.; Ruppenthal, S.; et al. Mutation of the Calmodulin Binding Motif IQ of the L-Type Cav1.2 Ca2+ Channel to EQ Induces Dilated Cardiomyopathy and Death. J. Biol. Chem. 2012, 287, 22616–22625. [Google Scholar] [CrossRef] [Green Version]
  131. Crotti, L.; Spazzolini, C.; Tester, D.J.; Ghidoni, A.; Baruteau, A.E.; Beckmann, B.M.; Behr, E.R.; Bennett, J.S.; Bezzina, C.R.; Bhuiyan, Z.A.; et al. Calmodulin Mutations and Life-Threatening Cardiac Arrhythmias: Insights from the International Calmodulinopathy Registry. Eur. Heart J. 2019, 40, 2964–2975. [Google Scholar] [CrossRef]
  132. Su, J.; Gao, Q.; Yu, L.; Sun, X.; Feng, R.; Shao, D.; Yuan, Y.; Zhu, Z.; Sun, X.; Kameyama, M.; et al. The LQT-Associated Calmodulin Mutant E141G Induces Disturbed Ca2+-Dependent Binding and a Flickering Gating Mode of the CaV1.2 Channel. Am. J. Physiol. Cell Physiol. 2020, 318, C991–C1004. [Google Scholar] [CrossRef] [Green Version]
  133. Minobe, E.; Mori, M.X.; Kameyama, M. Calmodulin and ATP Support Activity of the Cav1.2 Channel through Dynamic Interactions with the Channel. J. Physiol. 2017, 595, 2465–2477. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Crump, S.M.; Andres, D.A.; Sievert, G.; Satin, J. The Cardiac L-Type Calcium Channel Distal Carboxy Terminus Autoinhibition Is Regulated by Calcium. Am. J. Physiol. Heart Circ. Physiol. 2013, 304, H455–H464. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Lyu, L.; Gao, Q.; Xu, J.; Minobe, E.; Zhu, T.; Kameyama, M. A New Interaction between Proximal and Distal C-Terminus of Cav1.2 Channels. J. Pharmacol. Sci. 2017, 133, 240–246. [Google Scholar] [CrossRef] [PubMed]
  136. Lieb, A.; Ortner, N.; Striessnig, J. C-Terminal Modulatory Domain Controls Coupling of Voltage-Sensing to Pore Opening in Cav1.3 L-Type Ca2+ Channels. Biophys. J. 2014, 106, 1467–1475. [Google Scholar] [CrossRef] [Green Version]
  137. Scharinger, A.; Eckrich, S.; Vandael, D.H.; Schönig, K.; Koschak, A.; Hecker, D.; Kaur, G.; Lee, A.; Sah, A.; Bartsch, D.; et al. Cell-Type-Specific Tuning of Cav1.3 Ca2+-Channels by a C-Terminal Automodulatory Domain. Front. Cell Neurosci. 2015, 9, 309. [Google Scholar] [CrossRef] [Green Version]
  138. Kuzmenkina, E.; Novikova, E.; Jangsangthong, W.; Matthes, J.; Herzig, S. Single-Channel Resolution of the Interaction between C-Terminal CaV1.3 Isoforms and Calmodulin. Biophys. J. 2019, 116, 836–846. [Google Scholar] [CrossRef] [Green Version]
  139. Wahl-Schott, C.; Baumann, L.; Cuny, H.; Eckert, C.; Griessmeier, K.; Biel, M. Switching off Calcium-Dependent Inactivation in L-Type Calcium Channels by an Autoinhibitory Domain. Proc. Natl. Acad. Sci. USA 2006, 103, 15657–15662. [Google Scholar] [CrossRef] [Green Version]
  140. Singh, A.; Hamedinger, D.; Hoda, J.C.; Gebhart, M.; Koschak, A.; Romanin, C.; Striessnig, J. C-Terminal Modulator Controls Ca2+-Dependent Gating of Ca v1.4 L-Type Ca2+ Channels. Nat. Neurosci. 2006, 9, 1108–1116. [Google Scholar] [CrossRef]
  141. Yang, Y.; Yu, Z.; Geng, J.; Liu, M.; Liu, N.; Li, P.; Hong, W.; Yue, S.; Jiang, H.; Ge, H.; et al. Cytosolic Peptides Encoding CaV1 C-Termini Downregulate the Calcium Channel Activity-Neuritogenesis Coupling. Commun. Biol. 2022, 5, 484. [Google Scholar] [CrossRef]
  142. Singh, A.; Gebhart, M.; Fritsch, R.; Sinnegger-Brauns, M.J.; Poggiani, C.; Hoda, J.C.; Engel, J.; Romanin, C.; Striessnig, J.; Koschak, A. 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. 2008, 283, 20733–20744. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Striessnig, J.; Bolz, H.J.; Koschak, A. Channelopathies in Cav1.1, Cav1.3, and Cav1.4 Voltage-Gated L-Type Ca2+ Channels. Pflugers Arch. 2010, 460, 361–374. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Bock, G.; Gebhart, M.; Scharinger, A.; Jangsangthong, W.; Busquet, P.; Poggiani, C.; Sartori, S.; Mangoni, M.E.; Sinnegger-Brauns, M.J.; Herzig, S.; et al. Functional Properties of a Newly Identified C-Terminal Splice Variant of Cav1.3 L-Type Ca2+ Channels. J. Biol. Chem. 2011, 286, 42736–42748. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Liu, N.; Yang, Y.; Ge, L.; Liu, M.; Colecraft, H.M.; Liu, X. Cooperative and Acute Inhibition by Multiple C-Terminal Motifs of L-Type Ca2+ Channels. eLife 2017, 6, e21989. [Google Scholar] [CrossRef]
  146. Alseikhan, B.A.; DeMaria, C.D.; Colecraft, H.M.; Yue, D.T. Engineered Calmodulins Reveal the Unexpected Eminence of Ca2+ Channel Inactivation in Controlling Heart Excitation. Proc. Natl. Acad. Sci. USA 2002, 99, 17185–17190. [Google Scholar] [CrossRef] [Green Version]
  147. Liang, H.; DeMaria, C.D.; Erickson, M.G.; Mori, M.X.; Alseikhan, B.A.; Yue, D.T. Unified Mechanisms of Ca2+ Regulation across the Ca2+ Channel Family. Neuron 2003, 39, 951–960. [Google Scholar] [CrossRef] [Green Version]
  148. Tadross, M.R.; Dick, I.E.; Yue, D.T. Mechanism of Local and Global Ca2+ Sensing by Calmodulin in Complex with a Ca2+ Channel. Cell 2008, 133, 1228–1240. [Google Scholar] [CrossRef] [Green Version]
  149. Hye, Y.P.; Kim, S.A.; Korlach, J.; Rhoades, E.; Kwok, L.W.; Zipfel, W.R.; Waxham, M.N.; Webb, W.W.; Pollack, L. Conformational Changes of Calmodulin upon Ca2+ Binding Studied with a Microfluidic Mixer. Proc. Natl. Acad. Sci. USA 2008, 105, 542–547. [Google Scholar] [CrossRef] [Green Version]
  150. Limpitikul, W.B.; Greenstein, J.L.; Yue, D.T.; Dick, I.E.; Winslow, R.L. A Bilobal Model of Ca2+-Dependent Inactivation to Probe the Physiology of L-Type Ca2+ Channels. J. Gen. Physiol. 2018, 150, 1688–1701. [Google Scholar] [CrossRef] [Green Version]
  151. Hoang, J.; Prosser, R.S. Conformational Selection and Functional Dynamics of Calmodulin: A 19F Nuclear Magnetic Resonance Study. Biochemistry 2014, 53, 5727–5736. [Google Scholar] [CrossRef]
  152. Gsponer, J.; Christodoulou, J.; Cavalli, A.; Bui, J.M.; Richter, B.; Dobson, C.M.; Vendruscolo, M. A Coupled Equilibrium Shift Mechanism in Calmodulin-Mediated Signal Transduction. Structure 2008, 16, 736–746. [Google Scholar] [CrossRef] [PubMed]
  153. Westerlund, A.M.; Delemotte, L. Effect of Ca2+ on the Promiscuous Target-Protein Binding of Calmodulin. PLoS Comput. Biol. 2018, 14, e1006072. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Putkey, J.A.; Kleerekoper, Q.; Gaertner, T.R.; Waxham, M.N. A New Role for IQ Motif Proteins in Regulating Calmodulin Function. J. Biol. Chem. 2003, 278, 49667–49670. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Moradi, F.; Copeland, E.N.; Baranowski, R.W.; Scholey, A.E.; Stuart, J.A.; Fajardo, V.A. Calmodulin-Binding Proteins in Muscle: A Minireview on Nuclear Receptor Interacting Protein, Neurogranin, and Growth-Associated Protein 43. Int. J. Mol. Sci. 2020, 21, 1016. [Google Scholar] [CrossRef] [Green Version]
  156. Caprara, G.A.; Morabito, C.; Perni, S.; Navarra, R.; Guarnieri, S.; Mariggiò, M.A. Evidence for Altered Ca2+ Handling in Growth Associated Protein 43-Knockout Skeletal Muscle. Front. Physiol. 2016, 7, 493. [Google Scholar] [CrossRef] [Green Version]
  157. Pang, C.; Crump, S.M.; Jin, L.; Correll, R.N.; Finlin, B.S.; Satin, J.; Andres, D.A. Rem GTPase Interacts with the Proximal CaV1.2 C-Terminus and Modulates Calcium-Dependent Channel Inactivation. Channels 2010, 4, 192–202. [Google Scholar] [CrossRef] [Green Version]
  158. Tseng, P.Y.; Henderson, P.B.; Hergarden, A.C.; Patriarchi, T.; Coleman, A.M.; Lillya, M.W.; Montagut-Bordas, C.; Lee, B.; Hell, J.W.; Horne, M.C. α-Actinin Promotes Surface Localization and Current Density of the Ca2+ Channel CaV1.2 by Binding to the IQ Region of the A1 Subunit. Biochemistry 2017, 56, 3669–3681. [Google Scholar] [CrossRef]
  159. Haeseleer, F.; Sokal, I.; Verlinde, C.L.M.J.; Erdjument-Bromage, H.; Temps, P.; Pronin, A.N.; Benovic, J.L.; Fariss, R.N.; Palczewski, K. Five Members of a Novel Ca2+-Binding Protein (CABP) Subfamily with Similarity to Calmodulin. J. Biol. Chem. 2000, 275, 1247–1260. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  160. Zhou, H.; Yu, K.; McCoy, K.L.; Lee, A. Molecular Mechanism for Divergent Regulation of Cav1.2 Ca2+ Channels by Calmodulin and Ca2+-Binding Protein-1. J. Biol. Chem. 2005, 280, 29612–29619. [Google Scholar] [CrossRef] [Green Version]
  161. Findeisen, F.; Minor, D.L. Disruption of the IS6-AID Linker Affects Voltage-Gated Calcium Channel Inactivation and Facilitation. J. Gen. Physiol. 2009, 133, 327–343. [Google Scholar] [CrossRef] [Green Version]
  162. Findeisen, F.; Minor, D.L. Structural Basis for the Differential Effects of CaBP1 and Calmodulin on CaV1.2 Calcium-Dependent Inactivation. Structure 2010, 18, 1617–1631. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Findeisen, F.; Rumpf, C.H.; Minor, D.L. Apo States of Calmodulin and CaBP1 Control CaV1 Voltage-Gated Calcium Channel Function through Direct Competition for the IQ Domain. J. Mol. Biol. 2013, 425, 3217–3234. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Oz, S.; Benmocha, A.; Sasson, Y.; Sachyani, D.; Almagor, L.; Lee, A.; Hirsch, J.A.; Dascal, N. Competitive and Non-Competitive Regulation of Calcium-Dependent Inactivation in CaV1.2 L-Type Ca2+ Channels by Calmodulin and Ca2+-Binding Protein 1. J. Biol. Chem. 2013, 288, 12680–12691. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Yang, P.S.; Alseikhan, B.A.; Hiel, H.; Grant, L.; Mori, M.X.; Yang, W.; Fuchs, P.A.; Yue, D.T. Switching of Ca2+-Dependent Inactivation of Cav1.3 Channels by Calcium Binding Proteins of Auditory Hair Cells. J. Neurosci. 2006, 26, 10677–10689. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Cui, G.; Meyer, A.C.; Calin-Jageman, I.; Neef, J.; Haeseleer, F.; Moser, T.; Lee, A. Ca2+-Binding Proteins Tune Ca2+-Feedback to Cav1.3 Channels in Mouse Auditory Hair Cells. J. Physiol. 2007, 585, 791–803. [Google Scholar] [CrossRef] [PubMed]
  167. Ivanina, T.; Blumenstein, Y.; Shistik, E.; Barzilai, R.; Dascal, N. Modulation of L-Type Ca2+ Channels by Gβγ and Calmodulin via Interactions with N and C Termini of α1C. J. Biol. Chem. 2000, 275, 39846–39854. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  168. Pate, P.; Mochca-Morales, J.; Wu, Y.; Zhang, J.Z.; Rodney, G.G.; Serysheva, I.I.; Williams, B.Y.; Anderson, M.E.; Hamilton, S.L. Determinants for Calmodulin Binding on Voltage-Dependent Ca2+ Channels. J. Biol. Chem. 2000, 275, 39786–39792. [Google Scholar] [CrossRef] [Green Version]
  169. Tang, W.; Halling, D.B.; Black, D.J.; Pate, P.; Zhang, J.Z.; Pedersen, S.; Altschuld, R.A.; Hamilton, S.L. Apocalmodulin and Ca2+ Calmodulin-Binding Sites on the CaV1.2 Channel. Biophys. J. 2003, 85, 1538–1547. [Google Scholar] [CrossRef] [Green Version]
  170. Anderson, M.E. Ca2+-Dependent Regulation of Cardiac L-Type Ca2+ Channels: Is a Unifying Mechanism at Hand? J. Mol. Cell Cardiol. 2001, 33, 639–650. [Google Scholar] [CrossRef]
  171. Jurado, L.A.; Chockalingam, P.S.; Jarrett, H.W. Apocalmodulin. Physiol. Rev. 1999, 79, 661–682. [Google Scholar] [CrossRef]
  172. Saimi, Y.; Kung, C. Calmodulin as an Ion Channel Subunit. Annu. Rev. Physiol. 2002, 64, 289–311. [Google Scholar] [CrossRef] [PubMed]
  173. Gong, D.; Chi, X.; Wei, J.; Zhou, G.; Huang, G.; Zhang, L.; Wang, R.; Lei, J.; Chen, S.R.W.; Yan, N. Modulation of Cardiac Ryanodine Receptor 2 by Calmodulin. Nature 2019, 572, 347–351. [Google Scholar] [CrossRef] [PubMed]
  174. Kang, P.W.; Westerlund, A.M.; Shi, J.; White, K.M.F.; Dou, A.K.; Cui, A.H.; Silva, J.R.; Delemotte, L.; Cui, J. Calmodulin Acts as a State-Dependent Switch to Control a Cardiac Potassium Channel Opening. Sci. Adv. 2020, 6, eabd6798. [Google Scholar] [CrossRef]
  175. Romanin, C.; Gamsjaeger, R.; Kahr, H.; Schaufler, D.; Carlson, O.; Abernethy, D.R.; Soldatov, N.M. Ca2+ Sensors of L-Type Ca2+ Channel. FEBS Lett. 2000, 487, 301–306. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Erickson, M.G.; Alseikhan, B.A.; Peterson, B.Z.; Yue, D.T. Preassociation of Calmodulin with Voltage-Gated Ca2+ Channels Revealed by FRET in Single Living Cells. Neuron 2001, 31, 973–985. [Google Scholar] [CrossRef] [Green Version]
  177. Lian, L.Y.; Myatt, D.; Kitmitto, A. Apo Calmodulin Binding to the L-Type Voltage-Gated Calcium Channel Cav1.2 IQ Peptide. Biochem. Biophys. Res. Commun. 2007, 353, 565–570. [Google Scholar] [CrossRef]
  178. Black, D.J.; Halling, D.B.; Mandich, D.V.; Pedersen, S.E.; Altschuld, R.A.; Hamilton, S.L. Calmodulin Interactions with IQ Peptides from Voltage-Dependent Calcium Channels. Am. J. Physiol. Cell Physiol. 2005, 288, C669–C676. [Google Scholar] [CrossRef] [Green Version]
  179. Sun, W.; Feng, R.; Hu, H.; Guo, F.; Gao, Q.; Shao, D.; Yin, D.; Wang, H.; Sun, X.; Zhao, M.; et al. The Ca2+-Dependent Interaction of Calpastatin Domain L with the C-Terminal Tail of the Cav1.2 Channel. FEBS Lett. 2014, 588, 665–671. [Google Scholar] [CrossRef] [Green Version]
  180. Black, D.J.; Persechini, A. In Calmodulin-IQ Domain Complexes, the Ca2+-Free and Ca2+-Bound Forms of the Calmodulin C-Lobe Direct the N-Lobe to Different Binding Sites. Biochemistry 2011, 50, 10061–10068. [Google Scholar] [CrossRef]
  181. Liu, Y.; Yan, S.; Chen, S.; Zhang, J.; Shen, Y.; Su, J.; He, G.; Feng, R.; Shao, D.; Hao, L. Calmodulin Mutant in Central Linker Reduces the Binding Affinity with PreIQ and IQ While Interacting with CaV1.2 Channels. Biochem. Biophys. Res. Commun. 2020, 526, 78–84. [Google Scholar] [CrossRef]
  182. Asmara, H.; Minobe, E.; Saud, Z.A.; Kameyama, M. Interactions of Calmodulin with the Multiple Binding Sites of Cav1.2 Ca2+ Channels. J. Pharmacol. Sci. 2010, 112, 397–404. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Catterall, W.A.; Few, A.P. Calcium Channel Regulation and Presynaptic Plasticity. Neuron 2008, 59, 882–901. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  184. Liu, Z.; Vogel, H.J. Structural Basis for the Regulation of L-Type Voltage-Gated Calcium Channels: Interactions between the N-Terminal Cytoplasmic Domain and Ca2+-Calmodulin. Front. Mol. Neurosci. 2012, 5, 38. [Google Scholar] [CrossRef] [Green Version]
  185. Benmocha, A.; Almagor, L.; Oz, S.; Hirsch, J.A.; Dascal, N. Characterization of the Calmodulin-Binding Site in the N Terminus of CaV1.2. Channels 2009, 3, 337–342. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  186. Guggenheimer, A.B.; Almagor, L.; Tsemakhovich, V.; Tripathy, D.R.; Hirsch, J.A.; Dascal, N. Interactions between N and C Termini of α1C Subunit Regulate Inactivation of CaV1.2 L-Type Ca2+ Channel. Channels 2016, 10, 55–68. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  187. Gauberg, J.; Elkhatib, W.; Smith, C.L.; Singh, A.; Senatore, A. Divergent Ca2+/Calmodulin Feedback Regulation of CaV1 and CaV2 Voltage-Gated Calcium Channels Evolved in the Common Ancestor of Placozoa and Bilateria. J. Biol. Chem. 2022, 298. [Google Scholar] [CrossRef] [PubMed]
  188. Kawaji, K.; Minobe, E.; Mori, M.X.; Kameyama, M. Two Calmodulin Binding Site Model for the Regulation of Cav1.2 Channel. J. Physiol. Sci. 2016, 66, 76. [Google Scholar]
  189. Minobe, E.; Mori, M.X.; Kameyama, M. Cav1.2 Channel Inactivation Induced by Two Molecules of Calmodulin. J. Physiol. Sci. 2019, 69, 260. [Google Scholar]
  190. Ben-Johny, M.; Yang, P.S.; Bazzazi, H.; Yue, D.T. Dynamic Switching of Calmodulin Interactions Underlies Ca2+ Regulation of CaV1.3 Channels. Nat. Commun. 2013, 4, 1717. [Google Scholar] [CrossRef] [Green Version]
  191. Patton, D.E.; West, J.W.; Catterall, W.A.; Goldin, A.L. A Peptide Segment Critical for Sodium Channel Inactivation Functions as an Inactivation Gate in a Potassium Channel. Neuron 1993, 11, 967–974. [Google Scholar] [CrossRef]
  192. Shirokov, R.; Levis, R.; Shirokova, N.; Ríos, E. Ca2+-Dependent Inactivation of Cardiac L-Type Ca2+ Channels Does Not Affect Their Voltage Sensor. J. Gen. Physiol. 1993, 102, 1005–1030. [Google Scholar] [CrossRef] [Green Version]
  193. Zong, S.; Zhou, J.; Tanabe, T. Molecular Determinants of Calcium-Dependent Inactivation in Cardiac L-Type Calcium Channels. Biochem. Biophys. Res. Commun. 1994, 201, 1117–1123. [Google Scholar] [CrossRef] [PubMed]
  194. Imredy, J.P.; Yue, D.T. Mechanism of Ca2+-Sensitive Inactivation of L-Type Ca2+ Channels. Neuron 1994, 12, 1301–1318. [Google Scholar] [CrossRef] [PubMed]
  195. Cens, T.; Restituito, S.; Galas, S.; Charnet, P. Voltage and Calcium Use the Same Molecular Determinants to Inactivate Calcium Channels. J. Biol. Chem. 1999, 274, 5483–5490. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  196. Barrett, C.F.; Tsien, R.W. The Timothy Syndrome Mutation Differentially Affects Voltage- and Calcium-Dependent Inactivation of CaV1.2 L-Type Calcium Channels. Proc. Natl. Acad. Sci. USA 2008, 105, 2157–2162. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  197. Tadross, M.R.; Yue, D.T. Systematic Mapping of the State Dependence of Voltage- and Ca2+ dependent Inactivation Using Simple Open-Channel Measurements. J. Gen. Physiol. 2010, 135, 217–227. [Google Scholar] [CrossRef] [Green Version]
  198. Tadross, M.R.; Ben-Johny, M.; Yue, D.T. Molecular Endpoints of Ca2+/Calmodulin- and Voltage-Dependent Inactivation of Cav1.3 Channels. J. Gen. Physiol. 2010, 135, 197–215. [Google Scholar] [CrossRef] [Green Version]
  199. Raybaud, A.; Dodier, Y.; Bissonnette, P.; Simoes, M.; Bichet, D.G.; Sauvé, R.; Parent, L. The Role of the GX9GX3G Motif in the Gating of High Voltage-Activated Ca2+ Channels. J. Biol. Chem. 2006, 281, 39424–39436. [Google Scholar] [CrossRef] [Green Version]
  200. Almagor, L.; Chomsky-Hecht, O.; Ben-Mocha, A.; Hendin-Barak, D.; Dascal, N.; Hirsch, J.A. The Role of a Voltage-Dependent Ca2+ Channel Intracellular Linker: A Structure-Function Analysis. J. Neurosci. 2012, 32, 7602–7613. [Google Scholar] [CrossRef] [Green Version]
  201. Dick, I.E.; Joshi-Mukherjee, R.; Yang, W.; Yue, D.T. Arrhythmogenesis in Timothy Syndrome Is Associated with Defects in Ca2+-Dependent Inactivation. Nat. Commun. 2016, 7, 10370. [Google Scholar] [CrossRef] [Green Version]
  202. Limpitikul, W.B.; Dick, I.E.; Ben-Johny, M.; Yue, D.T. An Autism-Associated Mutation in CaV1.3 Channels Has Opposing Effects on Voltage- and Ca2+ -Dependent Regulation. Sci. Rep. 2016, 6, 27235. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  203. Abderemane-Ali, F.; Findeisen, F.; Rossen, N.D.; Minor, D.L. A Selectivity Filter Gate Controls Voltage-Gated Calcium Channel Calcium-Dependent Inactivation. Neuron 2019, 101, 1134–1149.e3. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  204. Babich, O.; Matveev, V.; Harris, A.L.; Shirokov, R. Ca2+-Dependent Inactivation of CaV1.2 Channels Prevents Gd3+ Block: Does Ca2+ Block the Pore of Inactivated Channels? J. Gen. Physiol. 2007, 129, 477–483. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  205. Cens, T.; Rousset, M.; Leyris, J.P.; Fesquet, P.; Charnet, P. Voltage- and Calcium-Dependent Inactivation in High Voltage-Gated Ca2+ Channels. Prog. Biophys. Mol. Biol. 2006, 90, 104–117. [Google Scholar] [CrossRef]
  206. Babich, O.; Isaev, D.; Shirokov, R. Role of Extracellular Ca2+ in Gating of CaV1.2 Channels. J. Physiol. 2005, 565, 709–715. [Google Scholar] [CrossRef]
  207. Wu, J.; Yan, Z.; Li, Z.; Yan, C.; Lu, S.; Dong, M.; Yan, N. Structure of the Voltage-Gated Calcium Channel Cav1.1 Complex. Science 2015, 350, aad2395. [Google Scholar] [CrossRef]
  208. Wu, J.; Yan, Z.; Li, Z.; Qian, X.; Lu, S.; Dong, M.; Zhou, Q.; Yan, N. Structure of the Voltage-Gated Calcium Channel Cav1.1 at 3.6 Å Resolution. Nature 2016, 537, 191–196. [Google Scholar] [CrossRef]
  209. Pitt, G.S. Calmodulin and CaMKII as Molecular Switches for Cardiac Ion Channels. Cardiovasc. Res. 2007, 73, 641–647. [Google Scholar] [CrossRef] [Green Version]
  210. Zhang, R.; Dzhura, I.; Grueter, C.E.; Thiel, W.; Colbran, R.J.; Anderson, M.E. A Dynamic α-β Inter-Subunit Agonist Signaling Complex Is a Novel Feedback Mechanism for Regulating L-Type Ca2+ Channel Opening. FASEB J. 2005, 19, 1573–1575. [Google Scholar] [CrossRef]
  211. Mori, M.X.; Erickson, M.G.; Yue, D.T. Functional Stoichiometry and Local Enrichment of Calmodulin Interacting with Ca2+ Channels. Science 2004, 304, 432–435. [Google Scholar] [CrossRef]
  212. Xiong, L.; Kleerekoper, Q.K.; He, R.; Putkey, J.A.; Hamilton, S.L. Sites on Calmodulin That Interact with the C-Terminal Tail of Cav1.2 Channel. J. Biol. Chem. 2005, 280, 7070–7079. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  213. Johny, M.B.; Yue, D.N.; Yue, D.T. A Novel FRET-Based Assay Reveals 1:1 Stoichiometry of Apocalmodulin Binding Across Voltage-Gated Ca and Na Ion Channels. Biophys. J. 2012, 102, 125a–126a. [Google Scholar] [CrossRef] [Green Version]
  214. Findeisen, F.; Minor, D.L. Progress in the Structural Understanding of Voltage-Gated Calcium Channel (CaV) Function and Modulation. Channels 2010, 4, 459–474. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  215. He, G.; Guo, F.; Zhu, T.; Shao, D.; Feng, R.; Yin, D.; Sun, X.; Hu, H.; Hwang, A.; Minobe, E.; et al. Lobe-Related Concentration- and Ca2+-Dependent Interactions of Calmodulin with C- and N-Terminal Tails of the CaV1.2 Channel. J. Physiol. Sci. 2013, 63, 345–353. [Google Scholar] [CrossRef]
  216. Shao, D.; Xu, J.; Zhao, M.; Sun, Y.; Minobe, E.; He, G.; Feng, R.; Hu, H.; Sun, X.; Guo, F.; et al. The Lobe-Specific Interaction of Calmodulin with the Cardiac CaV1.2 Channel. J. Physiol. Sci. 2014, 64, 177. [Google Scholar]
  217. Ben-Johny, M.; Yue, D.N.; Yue, D.T. Detecting Stoichiometry of Macromolecular Complexes in Live Cells Using FRET. Nat. Commun. 2016, 7, 13709. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  218. Chakouri, N.; Diaz, J.; Yang, P.S.; Ben-Johny, M. CaV Channels Reject Signaling from a Second CaM in Eliciting Ca2+-Dependent Feedback Regulation. J. Biol. Chem. 2020, 295, 14948–14962. [Google Scholar] [CrossRef]
  219. Fallon, J.L.; Halling, D.B.; Hamilton, S.L.; Quiocho, F.A. Structure of Calmodulin Bound to the Hydrophobic IQ Domain of the Cardiac Cav1.2 Calcium Channel. Structure 2005, 13, 1881–1886. [Google Scholar] [CrossRef] [Green Version]
  220. Navedo, M.F.; Cheng, E.P.; Yuan, C.; Votaw, S.; Molkentin, J.D.; Scott, J.D.; Santana, L.F. Increased Coupled Gating of L-Type Ca2+ Channels during Hypertension and Timothy Syndrome. Circ. Res. 2010, 106, 748–756. [Google Scholar] [CrossRef] [Green Version]
  221. Findeisen, F.; Tolia, A.; Arant, R.; Kim, E.Y.; Isacoff, E.Y.; Minor, D.L. Calmodulin Overexpression Does Not Alter Cav1.2 Function or Oligomerization State. Channels 2011, 5, 320–324. [Google Scholar] [CrossRef] [Green Version]
  222. Dixon, R.E.; Moreno, C.M.; Yuan, C.; Opitz-Araya, X.; Binder, M.D.; Navedo, M.F.; Santana, L.F. Graded Ca2+/Calmodulin-Dependent Coupling of Voltage-Gated CaV1.2 Channels. Elife 2015, 4, e05608. [Google Scholar] [CrossRef] [PubMed]
  223. Ito, D.W.; Hannigan, K.I.; Ghosh, D.; Xu, B.; Del Villar, S.G.; Xiang, Y.K.; Dickson, E.J.; Navedo, M.F.; Dixon, R.E. β-Adrenergic-Mediated Dynamic Augmentation of Sarcolemmal CaV1.2 Clustering and Co-Operativity in Ventricular Myocytes. J. Physiol. 2019, 597, 2139–2162. [Google Scholar] [CrossRef] [Green Version]
  224. Simms, B.A.; Zamponi, G.W. Trafficking and Stability of Voltage-Gated Calcium Channels. Cell. Mol. Life Sci. 2012, 69, 843–856. [Google Scholar] [CrossRef] [PubMed]
  225. Weiss, N.; Zamponi, G.W. Trafficking of Neuronal Calcium Channels. Neuronal Signal 2017, 1, NS20160003. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  226. Westhoff, M.; Dixon, R.E. Mechanisms and Regulation of Cardiac CaV1.2 Trafficking. Int. J. Mol. Sci 2021, 22, 5927. [Google Scholar] [CrossRef]
  227. Béguin, P.; Ng, Y.J.A.; Krause, C.; Mahalakshmi, R.N.; Mei, Y.N.; Hunziker, W. RGK Small GTP-Binding Proteins Interact with the Nucleotide Kinase Domain of Ca2+-Channel β-Subunits via an Uncommon Effector Binding Domain. J. Biol. Chem. 2007, 282, 11509–11520. [Google Scholar] [CrossRef] [Green Version]
  228. Béguin, P.; Mahalakshmi, R.N.; Nagashima, K.; Cher, D.H.K.; Kuwamura, N.; Yamada, Y.; Seino, Y.; Hunziker, W. Roles of 14-3-3 and Calmodulin Binding in Subcellular Localization and Function of the Small G-Protein Rem2. Biochem. J. 2005, 390, 67–75. [Google Scholar] [CrossRef]
  229. Béguin, P.; Mahalakshmi, R.N.; Nagashima, K.; Cher, D.H.K.; Takahashi, A.; Yamada, Y.; Seino, Y.; Hunziker, W. 14-3-3 and calmodulin control subcellular distribution of Kir/Gem and its regulation of cell shape and calcium channel activity. J. Cell Sci. 2005, 118 Pt 9, 1923–1934. [Google Scholar] [CrossRef] [Green Version]
  230. Béguin, P.; Mahalakshmi, R.N.; Nagashima, K.; Cher, D.H.K.; Ikeda, H.; Yamada, Y.; Seino, Y.; Hunziker, W. Nuclear Sequestration of β-Subunits by Rad and Rem Is Controlled by 14-3-3 and Calmodulin and Reveals a Novel Mechanism for Ca2+ Channel Regulation. J. Mol. Biol. 2006, 355, 34–46. [Google Scholar] [CrossRef]
  231. Béguin, P.; Nagashima, K.; Gonoi, T.; Shibasaki, T.; Takahashi, K.; Kashima, Y.; Ozaki, N.; Geering, K.; Iwanaga, T.; Seino, S. Regulation of Ca2+ Channel Expression at the Cell Surface by the Small G-Protein Kir/Gem. Nature 2001, 411, 701–706. [Google Scholar] [CrossRef]
  232. Wang, H.G.; George, M.S.; Kim, J.; Wang, C.; Pitt, G.S. Ca2+/Calmodulin Regulates Trafficking of CaV1.2 Ca2+ Channels in Cultured Hippocampal Neurons. J. Neurosci. 2007, 27, 9086–9093. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  233. Ravindran, A.; Qi, Z.L.; Harry, J.B.; Abrahimi, P.; Kobrinsky, E.; Soldatov, N.M. Calmodulin-Dependent Gating of Cav1.2 Calcium Channels in the Absence of Cavβ Subunits. Proc. Natl. Acad. Sci. USA 2008, 105, 8154–8159. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  234. Gao, T.; Bunemann, M.; Gerhardstein, B.L.; Ma, H.; Hosey, M.M. Role of the C Terminus of the α1C (CaV1.2) Subunit in Membrane Targeting of Cardiac L-Type Calcium Channels. J. Biol. Chem. 2000, 275, 25436–25444. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  235. Bourdin, B.; Marger, F.; Wall-Lacelle, S.; Schneider, T.; Klein, H.; Sauvé, R.; Parent, L. Molecular Determinants of the CaVβ-Induced Plasma Membrane Targeting of the CaV1.2 Channel. J. Biol. Chem. 2010, 285, 22853–22863. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  236. Hall, D.D.; Dai, S.; Tseng, P.Y.; Malik, Z.; Nguyen, M.; Matt, L.; Schnizler, K.; Shephard, A.; Mohapatra, D.P.; Tsuruta, F.; et al. Competition between α-Actinin and Ca2+-Calmodulin Controls Surface Retention of the L-Type Ca2+ Channel CaV1.2. Neuron 2013, 78, 483–497. [Google Scholar] [CrossRef] [Green Version]
  237. Persechini, A.; Stemmer, P.M. Calmodulin is a Limiting Factor in the Cell. Trends Cardiovasc. Med. 2002, 12, 32–37. [Google Scholar] [CrossRef]
Figure 1. Structures of calmodulin (CaM) and the Cav1.2 Ca2+ channel. (A,B) apoCaM ((A), PDB:1CFD) and Ca2+/CaM ((B), PDB:1UP5) shown in rainbow color from the N-terminus (blue) to C-terminus (red) with Ca2+ indicated by the yellow spheres. (C) Schematic structure of pore-forming α1C subunit of Cav1.2 channel. CaM binding sites are indicated in red and labeled as follows: NT, NSCaTE domain; I–II, a site in I–II loop; (A,C) regions of preIQ domain; and IQ, IQ domain. The C-terminal tail is divided into CT1, CT2, and CT3 regions.
Figure 1. Structures of calmodulin (CaM) and the Cav1.2 Ca2+ channel. (A,B) apoCaM ((A), PDB:1CFD) and Ca2+/CaM ((B), PDB:1UP5) shown in rainbow color from the N-terminus (blue) to C-terminus (red) with Ca2+ indicated by the yellow spheres. (C) Schematic structure of pore-forming α1C subunit of Cav1.2 channel. CaM binding sites are indicated in red and labeled as follows: NT, NSCaTE domain; I–II, a site in I–II loop; (A,C) regions of preIQ domain; and IQ, IQ domain. The C-terminal tail is divided into CT1, CT2, and CT3 regions.
Ijms 24 06409 g001
Figure 3. Models for CDI and CDF. (A) CDI model of Kim et al. [97]. Ca2+ binding to CaM, which is tethered to the preIQ-IQ region of the channel, induces a conformation change that initiates interaction with the EF-hand region, resulting in further interaction with the I–II loop, leading to CDI. (B) CDI model of Yue’s group [98]. Ca2+ binding to CaM induces bridging between the N- and C-termini, leading to CDI. (C) ‘Two CaM hypothesis’ [43,51,99]. At resting conditions, a fraction of channels are bound by CaM and the rest are CaM-free (low activity). (A) moderate increase in [Ca2+] increases the fraction of CaM-bound channels (high activity), leading to CDF. A further increase in [Ca2+] induces binding of the second Ca2+/CaM to the putative inactivation site, resulting in CDI. CaM’s N-lobe (light blue) and C-lobe (pink) with Ca2+-free (white circles) and partially Ca2+-saturated (pink circles) and Ca2+-saturated (red) EF-hand domains are illustrated. Other abbreviations are the same as those in Figure 1.
Figure 3. Models for CDI and CDF. (A) CDI model of Kim et al. [97]. Ca2+ binding to CaM, which is tethered to the preIQ-IQ region of the channel, induces a conformation change that initiates interaction with the EF-hand region, resulting in further interaction with the I–II loop, leading to CDI. (B) CDI model of Yue’s group [98]. Ca2+ binding to CaM induces bridging between the N- and C-termini, leading to CDI. (C) ‘Two CaM hypothesis’ [43,51,99]. At resting conditions, a fraction of channels are bound by CaM and the rest are CaM-free (low activity). (A) moderate increase in [Ca2+] increases the fraction of CaM-bound channels (high activity), leading to CDF. A further increase in [Ca2+] induces binding of the second Ca2+/CaM to the putative inactivation site, resulting in CDI. CaM’s N-lobe (light blue) and C-lobe (pink) with Ca2+-free (white circles) and partially Ca2+-saturated (pink circles) and Ca2+-saturated (red) EF-hand domains are illustrated. Other abbreviations are the same as those in Figure 1.
Ijms 24 06409 g003
Figure 4. Structure models of preIQ-IQ/Cav1.2 bound with CaM. (A) Crystallographic structure of two preIQ-IQ bound with four CaM. CaM N-lobes, tan; C-lobes, light blue; Ca2+, gold spheres. The preIQ domain (I and II) are blue to orange rainbow starting from the N termini (N); IQ domain of CaV I is red. The preIQ-IQ loop regions, IQ domain of CaV II and C-lobe of CaM IV are computer modeled (thin light gray lines). From [127] with permission. (B) Structure model of preIQ-IQ of Cav1.2 bound with two CaM. The ‘?’ signifies the potential that the second Ca2+/CaM anchors at the C region of preIQ to interact with other components of Cav1.2. Cavβ (green and blue) bound to I–II loop (red) are shown as a candidate of such components. Orientation of the C-terminal tail relative to the other intracellular components of Cav1.2 is not identified. From [107] with permission. For the second CaM, the C-lobe is bound to preIQ, and the N-lobe binds another region, possibly the Cavβ/I–II loop. (C,D). Structure model of preIQ-IQ of Cav1.2 bound with two CaM. The preIQ structure is assumed to bend or turn at 1610QANE1613 region (rabbit) between the A and C regions (C) to allow the second CaM to bridge these regions to form a tertiary complex (D). The R1601 and E1605, which interact with both the N− and C-lobes of CaM are also shown. The model is based on PDF:3G43. From [43] with permission.
Figure 4. Structure models of preIQ-IQ/Cav1.2 bound with CaM. (A) Crystallographic structure of two preIQ-IQ bound with four CaM. CaM N-lobes, tan; C-lobes, light blue; Ca2+, gold spheres. The preIQ domain (I and II) are blue to orange rainbow starting from the N termini (N); IQ domain of CaV I is red. The preIQ-IQ loop regions, IQ domain of CaV II and C-lobe of CaM IV are computer modeled (thin light gray lines). From [127] with permission. (B) Structure model of preIQ-IQ of Cav1.2 bound with two CaM. The ‘?’ signifies the potential that the second Ca2+/CaM anchors at the C region of preIQ to interact with other components of Cav1.2. Cavβ (green and blue) bound to I–II loop (red) are shown as a candidate of such components. Orientation of the C-terminal tail relative to the other intracellular components of Cav1.2 is not identified. From [107] with permission. For the second CaM, the C-lobe is bound to preIQ, and the N-lobe binds another region, possibly the Cavβ/I–II loop. (C,D). Structure model of preIQ-IQ of Cav1.2 bound with two CaM. The preIQ structure is assumed to bend or turn at 1610QANE1613 region (rabbit) between the A and C regions (C) to allow the second CaM to bridge these regions to form a tertiary complex (D). The R1601 and E1605, which interact with both the N− and C-lobes of CaM are also shown. The model is based on PDF:3G43. From [43] with permission.
Ijms 24 06409 g004
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Kameyama, M.; Minobe, E.; Shao, D.; Xu, J.; Gao, Q.; Hao, L. Regulation of Cardiac Cav1.2 Channels by Calmodulin. Int. J. Mol. Sci. 2023, 24, 6409. https://doi.org/10.3390/ijms24076409

AMA Style

Kameyama M, Minobe E, Shao D, Xu J, Gao Q, Hao L. Regulation of Cardiac Cav1.2 Channels by Calmodulin. International Journal of Molecular Sciences. 2023; 24(7):6409. https://doi.org/10.3390/ijms24076409

Chicago/Turabian Style

Kameyama, Masaki, Etsuko Minobe, Dongxue Shao, Jianjun Xu, Qinghua Gao, and Liying Hao. 2023. "Regulation of Cardiac Cav1.2 Channels by Calmodulin" International Journal of Molecular Sciences 24, no. 7: 6409. https://doi.org/10.3390/ijms24076409

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop