Skip to main content
Channels logoLink to Channels
. 2010 Nov 1;4(6):497–509. doi: 10.4161/chan.4.6.12871

G protein inhibition of CaV2 calcium channels

Kevin PM Currie 1,
PMCID: PMC3052249  PMID: 21150298

Abstract

Voltage-gated Ca2+ channels translate the electrical inputs of excitable cells into biochemical outputs by controlling influx of the ubiquitous second messenger Ca2+. As such the channels play pivotal roles in many cellular functions including the triggering of neurotransmitter and hormone release by CaV2.1 (P/Q-type) and CaV2.2 (N-type) channels. It is well established that G protein coupled receptors (GPCRs) orchestrate precise regulation neurotransmitter and hormone release through inhibition of CaV2 channels. Although the GPCRs recruit a number of different pathways, perhaps the most prominent, and certainly most studied among these is the so-called voltage-dependent inhibition mediated by direct binding of Gβγ to the α1 subunit of CaV2 channels. This article will review the basics of Ca2+-channels and G protein signaling, and the functional impact of this now classical inhibitory mechanism on channel function. It will also provide an update on more recent developments in the field, both related to functional effects and crosstalk with other signaling pathways, and advances made toward understanding the molecular interactions that underlie binding of Gβγ to the channel and the voltage-dependence that is a signature characteristic of this mechanism.

Key words: G protein, N-type, P/Q-type, CaV2, calcium channel, inhibition

Voltage-gated Ca2+ Channels

Voltage-gated calcium channels (Ca2+-channels) play pivotal roles in a diverse array of cellular functions, in large part by translating the electrical inputs of excitable cells into biochemical outputs due to influx of the ubiquitous second messenger Ca2+. Calcium triggers neurotransmitter and hormone release, muscle contraction, modulates ion channels to control excitability, the activity of many enzymes, gene expression, cellular differentiation and motility. Clearly there must be precise spatiotemporal regulation of calcium signaling in order to manifest these many downstream events in a controlled manner. In part this is accomplished by the rich functional diversity of Ca2+ channels mediated by distinct subunit combinations and complex modulation by a variety of second messengers and protein-protein interactions.

Initially, biophysical differences suggested the existence of multiple channel subtypes: low voltage-activated (LVA) channels require only modest membrane depolarization to open whereas high voltage-activated (HVA) channels require stronger depolarization. Differences in channel kinetics, and pharmacological sensitivity led to further distinction within the HVA class of channels: L-type channels were blocked by dihydropyridines, N-type channels blocked by ω-conotoxin GVIA, P/Q-type channels by ω-agatoxin IVA and R-type channels were resistant to these previously mentioned blockers but sensitive to SNX-482 (reviews of the history of Ca2+ channels3,4). To date, ten different mammalian genes that encode the pore forming α1 subunit of Ca2+ channels have been identified. Based on sequence homology these are grouped into three major families (reviewed in refs. 4 and 5): four CaV1 members (CaV1.1–CaV1.4; all L-type channels), three CaV2 members (CaV2.1, P/Q-type; CaV2.2, N-type; CaV2.3, R-type channels) and three CaV3 members (CaV3.1–CaV3.3, all T-type channels). The functional diversity of α1 subunits is even greater due to splice variants that can alter the biophysical properties, change the context and impact of other mutations, or even introduce novel regulatory sites.69

The α1-subunit consists of four homologous repeats (domains I–IV) each with six transmembrane spanning α-helical segments (S1–S6) and a ‘P-loop’ between S5 and S6 (Fig. 1). Each domain is reminiscent of a voltage-gated K+ channel subunit and it is thought that the channel adopts a similar tetrameric architecture with the S5–S6 and P-loop segments forming the pore domain and the S1–S4 segments (in particular the S4 helix that has multiple charged residues) forming the voltage-sensor domain. The intracellular N- and C-termini and the cytoplasmic loops connecting domains I–IV are all important for interaction with other proteins including the auxiliary β subunit of the channel, SNARE proteins (syntaxin, SNAP25), synaptotagmin, calmodulin and related Ca2+-binding proteins, various adapter proteins and kinases, and of note for this review, G protein βγ heterodimers (Gβγ) (Fig. 1).

Figure 1.

Figure 1

Schematic representation of voltage-gated Ca2+ channel topology and subunit composition. (A) Topology of the pore forming α1 subunit. There are four homologous repeats (domain I through domain IV) each with six transmembrane spanning α-helices (S1–S6) (blue or orange cylinders) and a ‘P-loop’ between S5 and S6. The pore domain (orange) comprises the S5–S6 and P-loop segments, and the voltage-sensor domain (blue) consists of the S1–S4 segments (in particular S4 that has multiple charged residues). The intracellular N- and C-termini and the cytoplasmic loops connecting domains I–IV are all important for interaction with other proteins including the auxiliary β subunit of the channel that binds to the AID on the I–II linker, synaptic proteins that interact at the “synprint site”, and G protein βγ heterodimers (Gβγ) that interact at three sites on the N-terminus, I–II linker and C-terminus (see text for more details). (B) Representation showing the 3D topology with the intracellular β subunit that interacts through its GK domain with the AID on the I–II linker of the α1 subunit. The α2δ subunit is largely extracellular and likely GPI-anchored to the plasma membrane.

High voltage-activated channels (CaV1 and CaV2) are hetero-oligomeric protein complexes and additionally contain a cytoplasmic β subunit (CaVβ) and a membrane associated α2δ subunit5 (Fig. 1). Four genes (each with multiple splice variants) encode CaVβ subunits (β1 through β4), and four genes encode α2δ subunits (α2δ-1 through α2δ-4). Both the α2δ and β subunits increase trafficking/plasma membrane expression of the α1 subunit and modulate channel kinetics. The α2δ subunit is the product of a single gene that undergoes posttranslational cleavage and subsequent formation of a disulphide bond between the α2 and δ components. The α2 component is extracellular and until recently it was considered that δ traversed the membrane giving rise to a type I transmembrane topology.10,11 However, in contrast to this assumption, recent evidence suggests that α2δ is glycerophosphatidylinositol (GPI)-anchored, and does not span the plasma membrane.12,13 Gabapentin and pregabalin, two widely prescribed drugs used for epilepsy, neuropathic pain and a variety of other off label uses, bind with high affinity to the α2δ-1 and α2δ-2 subunits, and thrombospondin was recently identified as an endogenous α2δ ligand.11,14 The CaVβ subunit has received considerable attention because of its prominent effects on channel trafficking and gating and particularly since the crystal structures of several subtypes were reported and shed light onto the structural elements that mediate these functional effects.1517 CaVβ is modular in nature with src-homology-3 (SH3) and guanylate kinase-like (GK) domains, placing it in the larger family of MAGUK proteins (reviewed in refs. 1820). The GK domain binds to the so-called α-interaction domain (AID) on the cytoplasmic I–II linker (Fig. 1). Notably, the AID partially overlaps with one of the binding sites for Gβγ heterodimers, and recent evidence suggests this CaVβ-AID interaction is critical for imparting voltage-dependence to Gβγ-mediated inhibition of CaV2 channels (discussed further below). Skeletal muscle (CaV1.1) channels also contain a γ-subunit but it remains unclear if there are direct interactions of γ with other CaV channel complexes.

The remainder of this review will focus on the CaV2 family of channels, in particular CaV2.1 (P/Q-type) and CaV2.2 (N-type) that are widely expressed in the mammalian nervous system and trigger neurotransmitter release at most synapses and also some neuroendocrine cells.

G Protein Coupled Receptors and Heterotrimeric G Proteins

Almost 800 human genes encode G protein coupled receptors (GPCRs) and over 30% of drugs on the market today target GPCRs.21,22 GPCRs are characterized by seven transmembrane spanning alpha helices with extracellular N-and intracellular C-termini. The receptors couple to heterotrimeric G proteins which act as molecular switches to transduce extracellular ligand binding into intracellular signaling cascades. In human, sixteen genes encode Gα subunits, five genes encode Gβ subunits and twelve genes encode Gγ subunits (reviewed in refs. 2326). In addition to transducin (Gαt) in the retina, there are four major families of Gα, and GPCRs are generally characterized according to the family of Gα subunit to which they couple: Gs coupled receptors typically activate adenylyl cyclase to elevate cAMP, Gi family receptors inhibit cAMP production, Gq family receptors activate phospholipase C signaling, and G12 receptor signaling is less well defined. Agonist binding to the GPCR catalyzes the exchange of GDP to GTP on the G protein α-subunit (Gα) causing conformational changes/dissociation of the Gα and Gβγ heterodimer.25,27 Subsequently, both the liberated Gα and Gβγ signal to multiple downstream effectors, including voltage-gated Ca2+-channels (see below). Intrinsic GTPase activity of the Gα subunit and subsequent reassociation of the Gα-GDP subunit with the βγ heterodimer terminates signaling. The GTPase activity can be accelerated by regulator of G protein signaling (RGS) proteins which consequently can influence the extent and duration of downstream events.28 Conversely, activator of G protein signaling (AGS) proteins can stimulate nucleotide exchange and lead to signaling independent of GPCR activation.29 Termination of signaling in the continued presence of agonist can also occur due to receptor desensitization. Desensitization is complex and for some receptors (e.g., the µ-opioid receptor) can be mediated by distinct signaling pathways depending upon the identity of the bound agonist.30 Phosphorylation by PKA, PKC or G protein coupled receptor kinases (GRKs) is involved and can result in uncoupling of the receptor from the downstream G proteins. Endocytic removal of the GPCR from the plasma membrane can also occur: GRKs, recruited by Gβγ, phosphorylate the C-terminus of the GPCR leading to recruitment of arrestins and subsequently the endocytic machinery.31,32 All of these mechanisms terminate G protein signaling and thus channel modulation, but in some cases, direct GPCR-CaV2 interactions have been demonstrated that result in endocytic removal of the channel from the plasma membrane upon GPCR desensitization.3335

Inhibition of Ca2+ Channels by G Protein βγ Subunits—Functional Effects

Since the first demonstration that norepinephrine could reduce action potential duration and Ca2+ current amplitude in chick sensory neurons,36,37 it has become clear that inhibition of Ca2+ channels by GPCRs is widespread and important for controlling neurotransmitter and hormone release (reviewed in refs. 3842). The diversity and complexity of GPCR/G protein signaling is huge and several distinct pathways converge on Ca2+ channels. Perhaps the most prominent, and certainly most studied among these is the so-called voltage-dependent inhibition mediated by direct binding of Gβγ to the α1 subunit of CaV2 channels. Voltage-independent inhibition is generally less well characterized and likely comprised of several distinct mechanisms including channel trafficking, phosphorylation and lipid signaling pathways (reviewed in refs. 34, 35, 40, 4348). The primary focus of this review is the direct, voltage-dependent inhibition of CaV2 channels.

Direct inhibition by GPCRs predominantly targets CaV2.2 (N-type) and CaV2.1 (P/Q-type) channels, although CaV2.3 channels are also inhibited by similar mechanisms. There are several characteristic features of the inhibition that have given rise to the voltage-dependent moniker: the peak amplitude of the whole-cell Ca2+-channel current (ICa) is reduced but this inhibition is diminished at depolarized membrane potentials; activation kinetics are slowed; the voltage-dependence of activation is shifted to more depolarized potentials; a conditioning prepulse to depolarized potentials can relieve most of the inhibition and normalize channel kinetics (termed prepulse relief or prepulse facilitation). These characteristics of the inhibition have been incorporated into models in which the channels exhibit two functional gating states, “willing” and “reluctant”.4953 The “reluctant” channels are bound to Gβγ and display the voltage-dependent shifts in channel gating noted above. Two isoforms of CaV2.2 were identified in rat brain and one of these, CaV2.2a, was constitutively reluctant in the absence of Gβγ due to a single amino acid substitution in domain IS3.54 Thus the reluctant gating state appears to be intrinsic to the channel and can be allosterically modulated by Gβγ, phosphorylation and CaVβ.55

Figure 2A shows an example of voltage-dependent inhibition of whole cell calcium current in adrenal chromaffin cells, a widely used neurosecretory model that is also useful to investigate autocrine/paracrine modulation of Ca2+ channels.56 The cells express P2Y purinergic receptors that are activated by ATP co-released with catecholamines.56,57 Prominent slowing of activation kinetics and prepulse relief of the inhibition is clearly seen. Such voltage-dependent relief of the inhibition is also physiologically relevant and can occur at least to some extent during high frequency trains of action potential-like waveforms (Fig. 2B).5863 In turn this might contribute to short term synaptic plasticity, at least at some synapses.64

Figure 2.

Figure 2

Functional effects of voltage-dependent inhibition. (A) current records from an adrenal chromaffin cell in the absence (black traces) and presence (red traces) of extracellular ATP. Chromaffin cells express P2Y receptors that are activated by ATP co-released with catecholamines. Hallmark features of voltage-dependent inhibition are shown and include reduced peak amplitude with prominent slowing of the activation kinetics (left pair of current traces). A conditioning prepulse to +100 mV applied just before the test pulse to +10 mV reversed most of the inhibition and normalized the activation kinetics of ICa. but had no effect under control conditions (black trace) (right pair of current traces) (B). Voltage-dependent relief of inhibition (facilitation) can also occur during stimuli that mimic brief trains of actions potentials. The example is taken from an adrenal chromaffin cell stimulated with a 45 Hz train of ten action potential-like stimuli (from −60 mV to +35 mV) in the presence of ATP to produce inhibition (red) or in control conditions (black traces). Currents were normalized to the first pulse of the train and mean data is shown on left. In control conditions there was a small decrease in current amplitude (probably due to inactivation), but in the presence of ATP there was a robust increase in current, reflecting partial reversal of the G protein mediated inhibition.

The general consensus is that voltage-dependent relief of the inhibition reflects a shift of the channels to the “willing” state due to transient dissociation of Gβγ (Fig. 3). Perhaps the most convincing evidence for this comes from kinetic analyses of prepulse relief as a function of agonist or Gβγ concentration. As the concentration of Gβγ was increased the rate of relief during the prepulse was not altered but the rate of reinhibition following the prepulse was accelerated.6568 This is the predicted outcome for voltage-dependent dissociation and rebinding of Gβγ. Furthermore, the kinetics of reinhibition were consistent with binding and unbinding of a single Gβγ dimer.68

Figure 3.

Figure 3

Voltage-dependent relief reflects transient dissociation of Gβγ from the channel at depolarized membrane potentials. Recombinant CaV2.2 channels expressed with β1b and α2δ in HEK293 cells. Gβγ was also co-expressed and produced tonic inhibition of ICa that could be reversed by a step to +100 mV. Lower part shows a series of superimposed currents. The interval between the conditioning prepulse to +100 mV and the following test pulse to +10 mV was increased from 10 ms to 200 ms. With longer intervals the increase in current amplitude becomes less as Gβγ rebinds to the channel. The blue line shows an exponential fit to the peak current amplitudes.

Single channel studies provided early evidence that the inhibition was “direct” and “membrane delimited”—in other words it did not involve a diffusible second messenger. In cell attached single channel recordings, bath application of agonist did not inhibit the channels whereas agonist in the patch pipette did.66,69,70 “Reluctant gating” has also been demonstrated in single channel recordings. Upon membrane depolarization, the delay before the channel first opens (latency) was increased during inhibition (i.e., the channels were “reluctant” to open), but there was little impact on other single channel parameters.51,71 Thus, the inhibited channels appeared essentially silenced, unable to open until Gβγ dissociated and the channels shifted to the “willing” state. However, two other studies reported that very brief channel openings from the “reluctant” state can occur in N-type channels (i.e., without Gβγ unbinding), although the probability of such events was low.52,53 In contrast, no such “reluctant openings” were observed for single P/Q-type channels. Overall, the dominant effects observed in all studies are the shift in activation and prolonged latency to first channel opening. The slow activation kinetics seen in whole cell recording (Fig. 2A) and delayed latency in single channel recordings reflect the conformational changes and subsequent dissociation of Gβγ from the channel upon membrane depolarization. Another functional manifestation of this voltagedependence is that rapid application of neurotransmitter during a depolarizing voltage-step produced little inhibition of whole cell ICa,72 presumably because binding of Gβγ is not favored by the conformational changes that accompany channel opening.

Shifts in activation gating are the dominant features of the “willing”-“reluctant” models and additional evidence for this comes from recordings of “gating currents” that reflect movement of the voltage-sensor domain of the channels in response to membrane potential changes. In recombinant CaV2.2 channels expressed in HEK cells these gating currents can be recorded in isolation as the cells lack endogenous voltage-gated channels. G proteins were found to reduce the amplitude and shift the voltage-dependence of gating currents to more depolarized potentials.73 G proteins also produced a significant separation in the voltagedependent activation of gating current and ionic current.73 These data are consistent with the idea that Gβγ binding slows movement of the voltage-sensor and uncouples or disrupts this movement from opening of the activation gate. More recently, gating current modulation by G proteins has also been reported in rat sympathetic neurons.74,75

Gβγ can also reduce inactivation of CaV2.2 channels,76,77 although this is somewhat more subtle and can be masked by concomitant voltage-dependent reversal of Gβγ-mediated inhibition (Gβγ unbinding). Inactivation of Ca2+ channels is complex and mediated by several voltage-dependent and Ca2+-dependent mechanisms.7880 The precise molecular correlates of inactivation remain somewhat unclear, but fast voltage-dependent inactivation might involve a “hinged lid” type mechanism, with the intracellular loop connecting domains I and II of the α1-subunit serving the role of the “inactivation gate” (reviewed in ref. 79, 81 and 82). The I-II loop is also an important binding site for Gβγ on the channel8386 (Fig. 1) (see below for more discussion). It is conceivable that binding of Gβγ disrupts movement of this putative inactivation gate, or its interaction with other channel domains, but this remains to be determined. It has also been proposed that CaV2.2 channels inactivate preferentially from intermediate closed state(s) favored during trains of brief depolarizing pulses.71 If Gβγ were to reduce the probability that the channels populate this state (from which inactivation is preferred) it might reduce the cumulative inactivation throughout a stimulus train.

Ca2+-dependent inactivation is mediated by calmodulin bound to the C-terminus of the channel α1 subunit.8791 It can be blocked in CaV2 channels, but not CaV1 channels, by strong intracellular Ca2+ buffering (EGTA or BAPTA in the patch pipette solution) indicating that it is mediated by a “global” elevation rather than a “local” microdomain of Ca2+. As such, the reduction of Ca2+-dependent inactivation by Gβγ76 might result from fewer channels opening and a diminished “global” Ca2+ signal. Consistent with this, blocking ∼50% of ICa by ω-conotoxin GVIA (i.e., by decreasing global Ca2+ entry) reduced calcium dependent inactivation, but a similar block of ICa by Cd2+ did not reduce inactivation.76 This highlights the complexity of Ca2+-dependent regulation of CaV channels and perhaps hints that the effects of Gβγ involve more than simply reducing global Ca2+ entry. A recent study proposed that allosteric hindrance of channel activation underlies Ca2+-dependent inactivation of CaV1.3,81 although it is not known if this holds true for CaV2 channels. Calmodulin might also facilitate interaction between the N and C terminal tails of Ca2+ channels,9294 and CaVβ binding to I–II linker appears to play a role in Ca2+-dependent inactivation.82 As discussed below, these regions of the channel are also intimately involved in voltage-dependent, Gβγ-mediated inhibition of CaV2 channels. Direct in vitro binding of Ca2+-calmodulin to Gβγ heterodimers has also been demonstrated.95 At present this is merely a list of coincident observations, but it is interesting to speculate that there might be crosstalk between Gβγ-mediated and calmodulin-mediated regulation of Ca2+ channels.

Differential Voltage-Dependent Inhibition within the CaV2 Family

Initially it was thought that CaV2.3 channels were not inhibited, but this depends on truncation of the α1 subunit N-terminus, whereas a splice variant with a longer N-terminus can be inhibited.9698 However, the majority of studies have focused on Nand P/Q-type channels in part because these are prominently inhibited and in part because they underlie the Ca2+ entry that triggers neurotransmitter release at most synapses and some neuroendocrine cells. Although very similar, there are subtle differences in the inhibition of CaV2.1 and CaV2.2 channels that might have important physiological consequences. The relative expression levels of the two types of channels differ between neurons and even between neighboring synapses arising from the same neuron. Thus, differential modulation of the two channels could lead to cell and/or synapse specific alterations in neuromodulation by GPCRs.

As already mentioned, differences in the inhibition of the two channels are apparent at the single channel level: very brief duration, low probability “reluctant openings” have been observed in CaV2.2 but not CaV2.1 channels.52,53 In whole cell recordings the peak amplitude of ICa is reduced substantially more for CaV2.2 than CaV2.1 ICa.99101 Moreover, trains of action potential-like stimuli relieve a greater proportion of CaV2.1 inhibition than CaV2.2 inhibition.102 These differences likely reflect the affinity of Gβγ binding to the channels. Indeed the apparent affinity of Gβγ (inferred from prepulse relief and reinhibition) is quite similar for the two channels at hyperpolarized or very depolarized potentials, but diverges at moderately depolarized levels (<+30 mV) within the physiologically relevant range of action potentials.52 Other evidence also suggests subtle differences in the binding of Gβγ heterodimers to CaV2.1 and CaV2.2 channels. For example, the rank order of inhibition produced by the five different Gβ subunits (Gβ1–5), all paired with the same Gγ2 subunit, differs for CaV2.1 and CaV2.2 channels.103 In another study, point mutations on the Gβ1 subunit had distinct effects on the inhibition of CaV2.1 and CaV2.2 channels.104

Functional differences might also arise from variable interactions or crosstalk with other signaling pathways. It has been shown that activation of PKC can reduce voltage-dependent inhibition of CaV2.2 channels (N-type) ICa.105109 Phosphorylation of Thr422 located in the I–II linker region of the rat α1 subunit underlies this effect and is close to one of the Gβγ binding sites.109,110 Of note, phosphorylation of Thr422 only disrupts inhibition mediated by Gβ1, but not other Gβ subunits.111 Thus the nature of the Gβγ heterodimer as well as the channel subunits will shape the functional impact of inhibition and its modulation by PKC. There are several other phosphorylation sites on CaV2.2 that might also contribute to channel regulation38 and there are also phosphorylation sites on CaVβ,20 although it is not known if these influence Gβγ-mediated inhibition. In some cases, the antagonistic effects of PKC might be mediated indirectly through phosphorylation of the GPCR.112

Inter- and Intra-Molecular Interactions that Mediate Direct Inhibition by Gβγ

Three distinct regions of CaV2 α1 subunits seem to participate in binding and/or transduction of the inhibition by Gβγ: the cytoplasmic linker between domains I and II (I–II linker), the N-terminus and the C-terminus (Fig. 1). Given that voltagedependent inhibition appears to involve a single Gβγ,68 these regions may contribute to a multi-site binding pocket.85 The C-terminus was identified as being essential for modulation of CaV2.3,113 but other studies suggest it plays more of a modulatory role in CaV2.2 channels, perhaps by increasing the affinity of binding and hence responses to low concentrations of free Gβγ.110,114,115 A variety of other signaling proteins bind to the C-terminus including calmodulin, CaMKII, PKC and Gα subunits.41,116 This might facilitate crosstalk with other signaling pathways to fine tune Gβγ-mediated inhibition, as suggested for binding of Gαq subunits and perhaps PKC.107,108 Splice variants of the CaV2.2 C-terminus have also been demonstrated to introduce a novel regulatory site for tyrosine kinases that underlies one form of voltage-independent inhibition.44

Studies investigating the role of the I–II linker in voltagedependent inhibition have at times come to somewhat differing conclusions about the relative importance of this region (reviewed in refs. 38 and 39). Two distinct binding sites have been identified within the I–II linker.83,84,109,117 The first partially overlaps with the binding site for the CaVβ subunit, the AID,16,118 and contains a QXXER consensus sequence for Gβγ found in other binding partners including phopholipase C β2 and type 2 adenylyl cyclase (QQIER in all three CaV2 family members). The second binding site for Gβγ identified downstream from the AID is termed the G protein interaction domain (GID).85 In vitro assays demonstrate high affinity (20–60 nM) interaction of Gβγ with the I–II linker,84,109,119 although partial masking of the first site by the CaVβ subunit reduces the affinity somewhat.120 Peptides based on both regions of the I–II linker diminish voltage-dependent inhibition of the channels and mutation to the sites can either reduce or enhance inhibition. Recently, a scanning mutagenesis approach systematically examined the involvement of each amino acid in these regions in voltage-dependent inhibition of rat CaV2.2 channels.117 Residues from 372–389 and 410–428 of the rat CaV2.2 channel were mutated to either alanine or cysteine (except for six residues in the 372–389 stretch shown to be buried by CaVβ binding). Two-out-of-thirty mutations (Arg376 and Val416 to alanine) significantly reduced the extent of voltage-dependent inhibition. Moreover, mutation of Arg376 to phenylalanine increased inhibition, further implicating this residue in binding of Gβγ and/or transduction of Gβγ binding into inhibition. It should be noted that some other studies have found a less prominent role for the I–II linker in voltage-dependent inhibition. For example, chimeric channels in which the I-II linker of CaV1.2 was introduced into the CaV2.2 backbone were still inhibited.100,121,122 However, recent studies would seem to support a central role for the I-II linker, at least for imparting voltage-dependence to the inhibition (discussed further below).120,123,124

The N-terminus has also emerged as essential in voltagedependent inhibition.121,122,125,126 Some of the first evidence for this came from the finding that CaV2.3 channels with a truncated N-terminus lacked inhibition, whereas the full-length N-terminus was permissive for inhibition.97 Likewise, truncating the N-terminal 55 amino acids of CaV2.2 prevented voltagedependent inhibition and introducing the CaV2.2 N-terminus into the CaV1.2 backbone conferred modest inhibition onto these normally resistant channels.97,121 The N-terminus has a glycine rich stretch of amino acids (9–40) that is followed by eleven amino acids (45–55) that are predicted to form an α-helix.127 This eleven amino acid stretch is critical for Gβγ-mediated inhibition of the channels, in particular residues S48, R52 and R54, with I49 also involved to a lesser extent.121 The functional importance of this region was confirmed and direct interaction with Gβγ demonstrated by the Yue group.122 Moreover, Agler et al.122 also demonstrated another role for the N-terminus: residues 56–95 directly bind to the N-terminal one-third of the I-II linker (from CaV2.2 but not CaV1.2). In the context of the full length channel this interaction is “gated” by Gβγ, and appears to be crucial for transducing the inhibition. Thus, the N-terminus contributes both to binding of Gβγ, and as an “inhibitory module” that through intra-molecular interaction with the I–II linker mediates the functional shift from willing to reluctant gating states. This finding might complicate interpretation of some studies investigating the role of the I–II linker. For example, point mutations within the proximal I–II linker region or peptides based on this sequence disrupt inhibition. This was interpreted as being due to disrupted binding of Gβγ to the I–II linker, but it is also possible that the perturbations disrupted binding of the N-terminus to the I–II linker. That being said, Gβγ can also bind to the I–II linker at the downstream GID site, and phosphorylation within this region disrupts Gβγ-mediated inhibition.110 This is consistent with disrupted binding of Gβγ to this region, but again, subtle allosteric effects that alter a somewhat distant binding site, or interaction of the upstream I–II linker with the N-terminus are also possible.

To summarize, current data suggest that the binding site for Gβγ is comprised from multiple sites on the N-terminus, I–II linker and probably C-terminus of the channel. Binding of Gβγ causes a conformational shift that facilitates interaction of the N-terminus (residues 56–95) with the initial one-third of the I–II-linker (Fig. 5). This (and perhaps other interactions) shifts gating charge movement to more depolarized potentials and uncouples voltage-sensor movement from channel activation.

Figure 5.

Figure 5

Model depicting the molecular interactions that underlie Gβγ-mediated inhibition of CaV2 channels. The pore forming α1 subunit is depicted as the blue cylinder with cytoplasmic domains illustrated by lines. (A) (upper three images) shows a channel with a CaVβ subunit (green oval) and (B) (lower three images) shows a channel without a CaVβ subunit. Current data suggest that the binding site for Gβγ (orange oval) is comprised from multiple sites on the Nterminus, I–II linker and probably C-terminus of the channel. Binding of Gβγ causes a conformational shift that facilitates interaction of the N-terminus “inhibitory module” (residues 56–95) with the initial one-third of the I-II-linker (middle parts-red arrow). This (and perhaps other interactions) shifts gating charge movement to more depolarized potentials and uncouples voltage-sensor movement from channel activation. Strong membrane depolarization (right parts) leads to conformational changes that result in unbinding of Gβγ and loss of interaction between the N-terminus and I–II linker. This depends upon binding of a CaVβ subunit to the AID on the I–II linker that induces a rigid α-helical connection to the upstream IS6 region of the pore and voltage-sensor. In the absence of CaVβ subunit binding, inhibition still occurs but cannot be reversed by strong depolarization (lower right part).

Structural Determinants on Gβγ

Several studies have also investigated the molecular/structural determinants on Gβγ that are responsible for interaction with CaV2 channels. Unlike the channel, there are high resolution crystal structures of Gβγ available both in isolation and bound to Gα, GRK2 and phosducin128132 (Fig. 4). Gβ adopts a toroidal, seven blade β-propeller structure with an α-helical N-terminal domain that binds to the α-helical N-terminus of the Gγ subunit. Gα interacts with multiple residues on the top face of Gβ and the side aspect of propeller blade 1 (Fig. 4). Many effectors bind to the same surface of Gβ that interacts with Gα and a protein interaction “hot spot” on this face has been identified with overlapping subsets of residues that are involved in binding to different effectors.133 Mutation of several residues on the Gα interacting surface disrupt inhibition of CaV2 channels,76,122,134,135 but residues on the opposite face of Gβ1 have also been implicated.135137 Moreover, Asn35 and Asn36 on Gβ1 have been shown to underlie the ability of PKC to antagonize inhibition of CaV2.2.136 Given that Thr422 on the rat CaV2.2 I–II linker has been identified as the phosphorylation site for PKC in this effect,110 it is tempting to speculate that these regions of the channel and Gβγ come into close proximity with one another. Figure 4 shows a molecular surface rendering of the Gαiβ1γ2 heterotrimer, and the Gβ1γ2 heterodimer based on the structure reported by Wall et al.128 Multiple residues that have been reported to disrupt inhibition of Ca2+ channels have been mapped onto this model. Most of these are on the same surface that Ga binds to, and are masked when Ga is present. There are also three residues on the reverse face of Gβ1.

Figure 4.

Figure 4

Molecular surface rendering of the heterotrimeric G protein structure (Gαi-green; Gβ1-cyan; and Gγ2-gray) based on structures reported by Wall et al.128 Molecular graphics images were produced using the UCSF Chimera package from the Resource for Biocomputing, Visualization and Informatics at the University of California, San Francisco (supported by NIH P41 RR001081).1,2 (A) shows a molecular surface rendering of the heterotrimer. The panels on the left and right are rotated ∼180° and show opposite faces of Gβγ. (B) shows the same images but with Gα subunit removed to reveal the underlying Gβ surface that is exposed upon G protein activation. Residues marked in red have been reported to disrupt inhibition of CaV2 channels. N35 and N36 (shown in magenta) are involved in crosstalk between Gβ1 and PKC phosphorylation of CaV2.2 (see text for details).

Another study reported that a peptide mimicking the N-terminal 25 amino acids of Gβ2 reduced inhibition of CaV2.1 (and activation of GIRK potassium channels).138 The Gβ N-terminal peptide disrupted FRET interaction between the Gβ2 and Gγ3 subunits suggesting a conformational shift or reorientation of the heterodimer that could disrupt interaction with the channels. Other studies have shown that the identity of Gγ within the Gβγ heterodimer can influence the extent of inhibition, with γ2 generally eliciting greater inhibition than γ1, γ3 or γ13.139,140 It remains unclear how Gγ exerts this influence, but it is interesting to note that the II–III linker of CaV2.1 contains a G-gamma-like (GGL) domain.141

Taken together, it is clear that multiple sites on both the CaV2 α1 subunit and the Gβγ heterodimer are involved in mediating voltage-dependent inhibition. However, the precise nature of these interactions remain unclear, including whether they represent direct protein-protein interactions or perhaps more subtle allosteric effects on the binding sites. Recently, small molecules and peptides that target the protein interaction “hot spot” on Gβγ have been shown to selectively disrupt interaction with some effectors (e.g., GRK2 and phospholipase Cβ).142 Some of these compounds have even demonstrated in vivo efficacy, for example to reduce the progression of heart failure through block of Gβ.-GRK2 interaction, or to reduce acute morphine induced anti-nociceptive tolerance perhaps through block of phospholipase Cβγ143,144 Thus, selective targeting of Gβγ-effector interactions is a viable proposition, so it is possible that compounds that enhance or disrupt modulation of calcium channels could be identified. Given that some point mutations on Gβ differentially altered inhibition of CaV2.1 and CaV2.2 channels,104 in principle it might even be possible to preferentially target the inhibition of one channel type over the other.

Why is the Inhibition Voltage-Dependent?

As already discussed, the consensus view is that voltage-dependent relief of the inhibition reflects transient dissociation of Gβγ from the channel at depolarized potentials. Recent data have identified some of the underlying molecular interactions that confer this voltage dependence to the inhibition. In particular, there appears to be an absolute requirement for CaVβ subunit binding to the AID in the I–II linker.120,123,124,145

The role of CaVβ subunits in voltage-dependent inhibition has been controversial with several seemingly opposing data sets and interpretations that have been reviewed in depth elsewhere.20,38,39 Briefly, one of the central questions that arose was whether or not CaVβ and Gβγ can both bind to the channel at the same time, or whether they compete for binding to the I–II linker in a mutually exclusive manner. Overlapping binding sites for the two proteins have been identified on the I–II linker, and initial observations showed that the functional effects of Gβγ generally oppose those of CaVβ. Contradictory FRET data suggested either competition146 or synergistic binding.147 Various other pieces of contradictory data were also obtained, and at least part of the explanation for this might stem from the presence of endogenous CaVβ subunits in some heterologous expression systems (including Xenopus oocytes) or confounding shifts in the voltage-dependence of activation by some CaVβ subunits (reviewed in ref. 38). Closer examination of functional effects at both the whole-cell and single-channel level strongly suggest that both proteins can and indeed must interact with the channel simultaneously to mediate voltage-dependent inhibition. Recently, several papers from the Dolphin and Yang labs have provided evidence that a major role for the CaVβ subunit is to confer voltage-dependence to Gβγ-mediated inhibition of CaV2.2,123,124,145 and CaV2.1 channels.120

The general approach used by both groups was to disrupt the CaVβ interaction with the AID to record currents from channels that lack a β subunit. The Dolphin lab mutated the AID of CaV2.2 channels (W391A) to reduce β subunit binding affinity by ∼1,000 fold.123 This resulted in a dramatic decrease in plasma membrane expression of the channels (due to disrupted trafficking) and alterations in gating kinetics indicating that the channels did indeed lack a β subunit. Activation of co-expressed D2 dopamine receptors still inhibited the mutant (W391A) channels to a similar degree as wild type channels, as did coexpression of Gβγ. However, prepulse reversal of the inhibition was almost abolished. Mutation of an additional two N-terminus residues shown to be essential for Gβγ-mediated inhibition (R52A and R54A-discussed above) abolished this voltage-independent inhibition in W391A channels. A follow up study demonstrated essentially the same findings when wild type CaV2.2 α1 subunit was expressed with α2δ but without CaVβ.124 Also in this study, the voltage-independent inhibition in the absence of the CaVβ was blocked by overexpression of transducin which acts to scavenge free Gβγ subunits. Thus, in the absence of CaVβ, inhibition of the channels by Gβγ was still present, but voltage-dependent reversal of the inhibition was abolished. The β1b subunit was used in the experiments outlined above, but when β2a was expressed with the W391A channels voltage-dependent relief of the Gβγ mediated inhibition was restored. Palmitoylation of two N-terminal cysteine residues promotes plasma membrane targeting of β2a, and mutation of these residues led to loss of voltage-dependent relief (i.e., the data resembled β1b). Thus, the authors proposed that palmitoylation enabled the local plasma membrane concentration of β2a to reach high enough levels that low affinity interaction with α1 could take place and permit voltage-dependent relief of the inhibition. In this context it is pertinent to note that previous work has shown the subtype of CaVβ can influence the extent and kinetics of Gβγ-mediated inhibition.148,149 Moreover, the particular effects of the CaVβ subunits are dependent on the identity of the G protein β subunit involved in producing the inhibition.149 Thus, the inhibition is precisely tuned by the identity of both the CaVβ and G protein β subunits.

Similar findings for CaV2.1 channels have also been reported recently in reference 120. In this study the authors used mutated CaVβ subunits that had reduced affinity for the AID. These were coexpressed in Xenopus oocytes with CaV2.1. Giant inside-out patches that contained many channels were pulled from the oocytes and macroscopic currents were recorded. Due to the reduced binding affinity of the mutant CaVβ subunit, washing the cytoplasmic face of the patches resulted in dissociation of CaVβ and this was confirmed by the expected shifts in channel kinetics compared to wild type. In the absence of CaVβ, application of purified Gβγ still inhibited the currents but prepulse reversal was abolished.

As outlined above, the CaVβ subunit has SH3 and GK domains that are separated by a variable HOOK region.18,20 The GK domain is solely responsible for binding to the AID on the I–II linker, although other regions including the SH3 and HOOK domains might interact elsewhere on the α1 subunit to modulate functional properties. For example, the HOOK domain has been reported to modulate inactivation kinetics of CaV2.2 channels.150 The Dolphin and Yang groups both demonstrated that binding of the isolated GK domain of CaVβ to the AID was sufficient to restore voltagedependence to the Gβγ-mediated inhibition120,124 (see below for further discussion). However, it is possible that the SH3 and variable HOOK domains of CaVβ might contribute in other ways. Deleting the HOOK domain resulted in tonic inhibition of CaV2.2 channels, perhaps due to increased affinity for the basal level of free Gβγ in the cells.124 Furthermore, at least some of the distinct effects of CaVβ2a on modulation are linked to its palmitoylation.124,149 In part this might be due to an effective concentration of the subunit at the plasma membrane, but other actions might also contribute. For example, the effects of CaVβ on channel kinetics are dependent on its orientation relative to the α1 subunit.151 It is possible that palmitoylation could alter the orientation of CaVβ to promote or hinder interactions between the SH3 or HOOK regions with as yet undetermined sites on the CaVα1 subunit. The palmitate groups on CaVβ2a have also been proposed to interact directly with the CaV2.2 subunit, and thereby mask a site for voltage-independent inhibition mediated by arachidonic acid.152,153

The molecular mechanism by which binding of CaVβ to the AID enables voltage-dependent relief of Gβγ mediated inhibition was further investigated by Zhang et al.120 In the absence of CaVβ the AID adopts a random coil but when CaVβ binds it induces the formation of an α-helical conformation that extends back to the interface with IS6.1517,154 Introducing seven glycines between the AID and IS6 to disrupt the α-helical structure prevented the ability of CaVβ to confer voltage-dependence to the inhibition. In contrast, introducing seven alanines (not expected to disrupt the α-helix) maintained the ability of CaVβ to confer voltage-dependence to Gβγ-mediated inhibition. Thus, it is possible that binding of CaVβ to the AID induces a rigid α-helical link with the upstream IS6 and this transmits movement of the voltage-sensor and activation gate (including IS6) to shift the I–II linker and alter the Gβγ binding pocket at depolarized potentials. It is also worth noting that Gβγ-mediated inhibition was still present (although slightly reduced) in both channel types lacking CaVβ, and in the CaV2.1 channels containing the seven glycine insert upstream of the AID.120,124 Thus, it would appear that the rigid α-helical link to the upstream activation gate and voltage-sensor is not required per se to transduce binding of Gβγ into functional inhibition.

Modulation of Gβγ-Mediated Inhibition by Other Proteins

Skeletal muscle L-type calcium channels (CaV1.1) have been shown to associate with a γ subunit in addition to a β and an α2δ subunit. Although several neuronal γ subunits have been identified that can alter expression of CaV2 channels, it remains uncertain that these proteins constitute bona fide channel subunits and the γ2 isoform (also called stargazin) and related proteins (γ3–7) associate with and modulate glutamatergic AMPA receptors.155 However, a recent study reported that stargazin binds Gβγ in vitro and that it scavenges Gβγ in Xenopus oocytes to reduce inhibition of CaV2.2 channels or activation of GIRK channels.156 Thus, the γ subunit, although not covalently bound to the channel complex, can modulate its properties by virtue of disrupting Gβγ-mediated inhibition.

CaV2 channels also interact with a variety of synaptic proteins.157,158 Syntaxin, SNAP25 and synaptotagmin all bind to the synaptic protein interaction site (synprint site) on the II–III linker of CaV2 channels (Fig. 1).159164 The functional consequences of these interactions are manyfold and reviewed in reference 157 and 158, but include targeting of the channels to synaptic release sites and modulation of channel kinetics. Syntaxin 1A (but not syntaxin 1B) enhanced voltage-dependent inhibition of CaV2 channels and in particular promotes tonic inhibition perhaps through a scaffolding role to colocalize the channels and G proteins.165167 Cleavage of syntaxin with botulinum neurotoxin C diminished inhibition of calcium channels in neuronal preparations supporting the notion that this interaction is physiologically important.168,169 Cysteine string protein also interacts with the synprint site and produces tonic inhibition of CaV2.2 channels perhaps due to colocalization of Gβγ and the channel and/or direct activation of GDP-GTP exchange (GEF activity) on the Gα subunit to liberate Gβγ.170,171

In recent years considerable attention has been given to regulator of G protein signaling (RGS) proteins which accelerate the intrinsic GTPase activity of Gα subunits. There are over 20 members of the RGS family and several of these modulate G protein inhibition of calcium channels.28 In part this involves the well documented acceleration of intrinsic GTPase activity to limit the availability of Gβγ. However, RGS proteins have several other modes of action172 including some that dimerize with Gβ5 subunits and some (e.g., RGS12) that directly interact with CaV2.2 channels.86,173 Experimental manipulation of RGS expression has been demonstrated to alter calcium channel inhibition174,175 and it has also been shown that this in turn can alter short term synaptic plasticity.176 Expression of several RGS proteins can be dynamically regulated by physiological and pathophysiological situations. For example, dopamine depletion (as occurs in Parkinson's disease) upregulated RGS4 in striatal interneurons resulting in diminished autoinhibition of calcium channels by M4 muscarinic receptors.177 Thus, RGS proteins have emerged as important modulators of Gβγ-mediated inhibition (among other pathways) and there is considerable interest in targeting these proteins for therapeutic development.178

Conversely, activator of G protein signaling (AGS) proteins can stimulate nucleotide exchange and lead to signaling independent of GPCR activation.29 Coexpression of AGS1 or the related Rhes protein with CaV2.2 in HEK293 cells induced tonic inhibition due to liberation of Gβγ from Gi/o-type heterotrimers.179 Thus it is possible that such signaling can occur, but this remains to be demonstrated in native channels under a physiological (or pathophysiological) context.

As outlined above, even though the voltage-dependent inhibition is mediated by Gβγ, in most cases the GPCR couples to pertussis toxin sensitive Gi/o-type G proteins. This functional selectivity might be achieved in several ways. For example, Gq-coupled receptors have been shown to couple (although not exclusively) with Gβ5 which in general is less efficacious at producing inhibition. The Gαq subunit might also associate directly with the channel to blunt the inhibition through PKC mediated phosphorylation.107,108 Voltage-dependent inhibition is membrane delimited so the GPCR and channel must be in close proximity. In part this might be mediated through adapter proteins like Homer180 or NHERF2.181 Direct interaction of GPCRs and CaV2.2 channels has also been demonstrated for opioid-like receptors (ORL1). Constitutive activity of the receptors leads to tonic voltage-dependent inhibition, and upon prolonged agonist application the channel is internalized as a consequence of desensitization of the coupled ORL1 receptor.35,182 More recently it has been shown that ORL1 and classic opioid receptors (e.g., µ-opioid receptors) can heterodimerize. One notable consequence of this is that µ-opioid receptor agonists lead to internalization of the GPCR-channel complex, which does not occur in the absence of the ORL1 receptor.183 D1 and D2 dopamine receptors also bind directly to CaV2.2 channels and modulate channel trafficking33,184 and GABAB receptors also lead to channel internalization although the mechanism is somewhat different.34

It is also worth noting that point mutations linked to familial hemiplegic migraine (FHM mutations) have been identified in CaV2.1 and lead to complex alterations in channel function.185,186 Many of these seem to exert “gain-of-function” effects, including three (R192Q, S218L, Y1245C) that diminish Gβγ-mediated inhibition.187189

Overall, it is clear that a multitude of protein-protein interactions and second messengers can fine tune the inhibition of CaV2 channels by G proteins. Membrane potential, firing patterns, channel subunit composition/splice variants and Gβγ heterodimer composition all modulate the extent and/or kinetics of inhibition. Indeed, the complexity of calcium channel inhibition is even greater because there are also several mechanisms leading to voltage-independent inhibition, including multiple kinases, phosphatases and lipid signaling pathways. These mechanisms are outside the scope of this review, but it is worth mentioning that some also involve Gβγ.190 Direct (but voltage-independent) inhibition of CaV3.2, T-type channels by Gβγ has also been reported.191,192 Ultimately, this complexity is linked to the overriding need to precisely control the calcium entry that plays such pivotal roles in cellular physiology, including the triggering and modulation of neurotransmitter release.

Acknowledgements

I would like to thank Mark Jewell (currently a graduate student in my lab) for help with the molecular graphics images in Figure 4 that were produced using the UCSF Chimera package from the Resource for Biocomputing, Visualization and Informatics at the University of California, San Francisco (supported by NIH P41 RR001081).1,2 Work in my lab is supported by the National Institutes of Health National Institute Of Neurological Disorders And Stroke [Grant R01-NS052446].

References

  • 1.Pettersen EF, Goddard TD, Huang CC, Couch GS, Greenblatt DM, Meng EC, et al. UCSF Chimera—a visualization system for exploratory research and analysis. J Comput Chem. 2004;25:1605–1612. doi: 10.1002/jcc.20084. [DOI] [PubMed] [Google Scholar]
  • 2.Sanner MF, Olson AJ, Spehner JC. Reduced surface: an efficient way to compute molecular surfaces. Biopolymers. 1996;38:305–320. doi: 10.1002/(SICI)1097-0282(199603)38:3%3C305::AID-BIP4%3E3.0.CO;2-Y. [DOI] [PubMed] [Google Scholar]
  • 3.Barrett CF, Tsien RW. Brief history of calcium channel discovery. In: Zamponi GW, editor. Voltage gated calcium channels. Plenum: Kluwer Academic; 2004. pp. 1–21. [Google Scholar]
  • 4.Dolphin AC. A short history of voltage-gated calcium channels. Br J Pharmacol. 2006;147:56–62. doi: 10.1038/sj.bjp.0706442. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Catterall WA. Structure and regulation of voltage-gated Ca2+ channels. Annu Rev Cell Dev Biol. 2000;16:521–555. doi: 10.1146/annurev.cellbio.16.1.521. [DOI] [PubMed] [Google Scholar]
  • 6.Gray AC, Raingo J, Lipscombe D. Neuronal calcium channels: splicing for optimal performance. Cell Calcium. 2007;42:409–417. doi: 10.1016/j.ceca.2007.04.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Liao P, Zhang HY, Soong TW. Alternative splicing of voltage-gated calcium channels: from molecular biology to disease. Pflugers Arch. 2009;458:481–487. doi: 10.1007/s00424-009-0635-5. [DOI] [PubMed] [Google Scholar]
  • 8.Lipscombe D, Raingo J. Alternative splicing matters: N-type calcium channels in nociceptors. Channels (Austin) 2007;1:225–227. doi: 10.4161/chan.4809. [DOI] [PubMed] [Google Scholar]
  • 9.Adams PJ, Garcia E, David LS, Mulatz KJ, Spacey SD, Snutch TP. CaV2.1 P/Q-type calcium channel alternative splicing affects the functional impact of familial hemiplegic migraine mutations: implications for calcium channelopathies. Channels (Austin) 2009;3:110–121. doi: 10.4161/chan.3.2.7932. [DOI] [PubMed] [Google Scholar]
  • 10.Arikkath J, Campbell KP. Auxiliary subunits: essential components of the voltage-gated calcium channel complex. Curr Opin Neurobiol. 2003;13:298–307. doi: 10.1016/s0959-4388(03)00066-7. [DOI] [PubMed] [Google Scholar]
  • 11.Davies A, Hendrich J, Van Minh AT, Wratten J, Douglas L, Dolphin AC. Functional biology of the α2δ subunits of voltage-gated calcium channels. Trends Pharmacol Sci. 2007;28:220–228. doi: 10.1016/j.tips.2007.03.005. [DOI] [PubMed] [Google Scholar]
  • 12.Bauer CS, Tran-Van-Minh A, Kadurin I, Dolphin AC. A new look at calcium channel α2δ subunits. Curr Opin Neurobiol. 2010;20:563–571. doi: 10.1016/j.conb.2010.05.007. [DOI] [PubMed] [Google Scholar]
  • 13.Davies A, Kadurin I, Alvarez-Laviada A, Douglas L, Nieto-Rostro M, Bauer CS, et al. The α2δ subunits of voltage-gated calcium channels form GPI-anchored proteins, a posttranslational modification essential for function. Proc Natl Acad Sci USA. 2010;107:1654–1659. doi: 10.1073/pnas.0908735107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Eroglu C, Allen NJ, Susman MW, O'Rourke NA, Park CY, Ozkan E, et al. Gabapentin receptor α2δ-1 is a neuronal thrombospondin receptor responsible for excitatory CNS synaptogenesis. Cell. 2009;139:380–392. doi: 10.1016/j.cell.2009.09.025. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Opatowsky Y, Chen CC, Campbell KP, Hirsch JA. Structural analysis of the voltage-dependent calcium channel β subunit functional core and its complex with the α1 interaction domain. Neuron. 2004;42:387–399. doi: 10.1016/s0896-6273(04)00250-8. [DOI] [PubMed] [Google Scholar]
  • 16.Van Petegem F, Clark KA, Chatelain FC, Minor DL., Jr Structure of a complex between a voltage-gated calcium channel β-subunit and an α-subunit domain. Nature. 2004;429:671–675. doi: 10.1038/nature02588. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Chen YH, Li MH, Zhang Y, He LL, Yamada Y, Fitzmaurice A, et al. Structural basis of the α1-β subunit interaction of voltage-gated Ca2+ channels. Nature. 2004;429:675–680. doi: 10.1038/nature02641. [DOI] [PubMed] [Google Scholar]
  • 18.Dolphin AC. B subunits of voltage-gated calcium channels. J Bioenerg Biomembr. 2003;35:599–620. doi: 10.1023/b:jobb.0000008026.37790.5a. [DOI] [PubMed] [Google Scholar]
  • 19.Hidalgo P, Neely A. Multiplicity of protein interactions and functions of the voltage-gated calcium channel β-subunit. Cell Calcium. 2007;42:389–396. doi: 10.1016/j.ceca.2007.05.009. [DOI] [PubMed] [Google Scholar]
  • 20.Buraei Z, Yang J. The β subunit of voltage-gated Ca2+ channels. Physiol Rev. 2010;90:1461–1506. doi: 10.1152/physrev.00057.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Bjarnadottir TK, Gloriam DE, Hellstrand SH, Kristiansson H, Fredriksson R, Schioth HB. Comprehensive repertoire and phylogenetic analysis of the G protein-coupled receptors in human and mouse. Genomics. 2006;88:263–273. doi: 10.1016/j.ygeno.2006.04.001. [DOI] [PubMed] [Google Scholar]
  • 22.Hopkins AL, Groom CR. The druggable genome. Nat Rev Drug Discov. 2002;1:727–730. doi: 10.1038/nrd892. [DOI] [PubMed] [Google Scholar]
  • 23.Kristiansen K. Molecular mechanisms of ligand binding, signaling and regulation within the superfamily of G-protein-coupled receptors: molecular modeling and mutagenesis approaches to receptor structure and function. Pharmacol Ther. 2004;103:21–80. doi: 10.1016/j.pharmthera.2004.05.002. [DOI] [PubMed] [Google Scholar]
  • 24.Downes GB, Gautam N. The G protein subunit gene families. Genomics. 1999;62:544–552. doi: 10.1006/geno.1999.5992. [DOI] [PubMed] [Google Scholar]
  • 25.Oldham WM, Hamm HE. Heterotrimeric G protein activation by G-protein-coupled receptors. Nat Rev Mol Cell Biol. 2008;9:60–71. doi: 10.1038/nrm2299. [DOI] [PubMed] [Google Scholar]
  • 26.Rosenbaum DM, Rasmussen SG, Kobilka BK. The structure and function of G-protein-coupled receptors. Nature. 2009;459:356–363. doi: 10.1038/nature08144. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.McIntire WE. Structural determinants involved in the formation and activation of G protein βγ dimers. Neurosignals. 2009;17:82–99. doi: 10.1159/000186692. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Hollinger S, Hepler JR. Cellular regulation of RGS proteins: modulators and integrators of G protein signaling. Pharmacol Rev. 2002;54:527–559. doi: 10.1124/pr.54.3.527. [DOI] [PubMed] [Google Scholar]
  • 29.Sato M, Blumer JB, Simon V, Lanier SM. Accessory proteins for G proteins: partners in signaling. Ann Rev Pharmacol Toxicol. 2006;46:151–187. doi: 10.1146/annurev.pharmtox.46.120604.141115. [DOI] [PubMed] [Google Scholar]
  • 30.Kelly E, Bailey CP, Henderson G. Agonist-selective mechanisms of GPCR desensitization. Br J Pharmacol. 2008;153:379–388. doi: 10.1038/sj.bjp.0707604. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Ferguson SS. Evolving concepts in G protein-coupled receptor endocytosis: the role in receptor desensitization and signaling. Pharmacol Rev. 2001;53:1–24. [PubMed] [Google Scholar]
  • 32.Gainetdinov RR, Premont RT, Bohn LM, Lefkowitz RJ, Caron MG. Desensitization of G protein-coupled receptors and neuronal functions. Annu Rev Neurosci. 2004;27:107–144. doi: 10.1146/annurev.neuro.27.070203.144206. [DOI] [PubMed] [Google Scholar]
  • 33.Kisilevsky AE, Mulligan SJ, Altier C, Iftinca MC, Varela D, Tai C, et al. D1 receptors physically interact with N-type calcium channels to regulate channel distribution and dendritic calcium entry. Neuron. 2008;58:557–570. doi: 10.1016/j.neuron.2008.03.002. [DOI] [PubMed] [Google Scholar]
  • 34.Tombler E, Cabanilla NJ, Carman P, Permaul N, Hall JJ, Richman RW, et al. G protein-induced trafficking of voltage-dependent calcium channels. J Biol Chem. 2006;281:1827–1839. doi: 10.1074/jbc.M508829200. [DOI] [PubMed] [Google Scholar]
  • 35.Altier C, Khosravani H, Evans RM, Hameed S, Peloquin JB, Vartian BA, et al. ORL1 receptor-mediated internalization of N-type calcium channels. Nat Neurosci. 2006;9:31–40. doi: 10.1038/nn1605. [DOI] [PubMed] [Google Scholar]
  • 36.Dunlap K, Fischbach GD. Neurotransmitters decrease the calcium ocmponent of sensory neurone action potentials. Nature. 1978;276:837–839. doi: 10.1038/276837a0. [DOI] [PubMed] [Google Scholar]
  • 37.Dunlap K, Fischbach GD. Neurotransmitters decrease the calcium conductance activated by depolarization of embryonic chick sensory neurones. J Physiol. 1981;317:519–535. doi: 10.1113/jphysiol.1981.sp013841. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Dolphin AC. G protein modulation of voltage-gated calcium channels. Pharmacol Rev. 2003;55:607–627. doi: 10.1124/pr.55.4.3. [DOI] [PubMed] [Google Scholar]
  • 39.Tedford HW, Zamponi GW. Direct G protein modulation of CaV2 calcium channels. Pharmacol Rev. 2006;58:837–862. doi: 10.1124/pr.58.4.11. [DOI] [PubMed] [Google Scholar]
  • 40.Ikeda SR, Dunlap K. Voltage-dependent modulation of N-type calcium channels: role of G protein subunits. Adv Second Messenger Phosphoprotein Res. 1999;33:131–151. doi: 10.1016/s1040-7952(99)80008-1. [DOI] [PubMed] [Google Scholar]
  • 41.Catterall WA, Few AP. Calcium channel regulation and presynaptic plasticity. Neuron. 2008;59:882–901. doi: 10.1016/j.neuron.2008.09.005. [DOI] [PubMed] [Google Scholar]
  • 42.Stephens GJ. G-protein-coupled-receptor-mediated presynaptic inhibition in the cerebellum. Trends Pharmacol Sci. 2009;30:421–430. doi: 10.1016/j.tips.2009.05.008. [DOI] [PubMed] [Google Scholar]
  • 43.Suh BC, Leal K, Hille B. Modulation of high-voltage activated Ca2+ channels by membrane phosphatidylinositol 4,5-bisphosphate. Neuron. 2010;67:224–238. doi: 10.1016/j.neuron.2010.07.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Raingo J, Castiglioni AJ, Lipscombe D. Alternative splicing controls G protein-dependent inhibition of N-type calcium channels in nociceptors. Nat Neurosci. 2007;10:285–292. doi: 10.1038/nn1848. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Liu L, Heneghan JF, Michael GJ, Stanish LF, Egertova M, Rittenhouse AR. L- and N-current but not M-current inhibition by M1 muscarinic receptors requires DAG lipase activity. J Cell Physiol. 2008;216:91–100. doi: 10.1002/jcp.21378. [DOI] [PubMed] [Google Scholar]
  • 46.Hille B. Modulation of ion-channel function by G-protein-coupled receptors. Trends Neurosci. 1994;17:531–6. doi: 10.1016/0166-2236(94)90157-0. [DOI] [PubMed] [Google Scholar]
  • 47.Elmslie KS. Neurotransmitter modulation of neuronal calcium channels. J Bioenerg Biomembr. 2003;35:477–489. doi: 10.1023/b:jobb.0000008021.55853.18. [DOI] [PubMed] [Google Scholar]
  • 48.Roberts-Crowley ML, Mitra-Ganguli T, Liu L, Rittenhouse AR. Regulation of voltage-gated Ca2+ channels by lipids. Cell Calcium. 2009;45:589–601. doi: 10.1016/j.ceca.2009.03.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Bean BP. Neurotransmitter inhibition of neuronal calcium currents by changes in channel voltage dependence. Nature. 1989;340:153–156. doi: 10.1038/340153a0. [DOI] [PubMed] [Google Scholar]
  • 50.Elmslie KS, Zhou W, Jones SW. LHRH and GTP-γ-S modify calcium current activation in bullfrog sympathetic neurons. Neuron. 1990;5:75–80. doi: 10.1016/0896-6273(90)90035-e. [DOI] [PubMed] [Google Scholar]
  • 51.Carabelli V, Lovallo M, Magnelli V, Zucker H, Carbone E. Voltage-dependent modulation of single N-Type Ca2+ channel kinetics by receptor agonists in IMR32 cells. Biophys J. 1996;70:2144–2154. doi: 10.1016/S0006-3495(96)79780-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Colecraft HM, Patil PG, Yue DT. Differential occurrence of reluctant openings in G-protein-inhibited N- and P/Q-type calcium channels. J Gen Physiol. 2000;115:175–192. doi: 10.1085/jgp.115.2.175. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Lee HK, Elmslie KS. Reluctant gating of single N-type calcium channels during neurotransmitter-induced inhibition in bullfrog sympathetic neurons. J Neurosci. 2000;20:3115–3128. doi: 10.1523/JNEUROSCI.20-09-03115.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Zhong H, Li B, Scheuer T, Catterall WA. Control of gating mode by a single amino acid residue in transmembrane segment IS3 of the N-type Ca2+ channel. Proc Natl Acad Sci USA. 2001;98:4705–4709. doi: 10.1073/pnas.051629098. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Herlitze S, Zhong H, Scheuer T, Catterall WA. Allosteric modulation of Ca2+ channels by G proteins, voltage-dependent facilitation, protein kinase C and CaVb subunits. Proc Natl Acad Sci USA. 2001;98:4699–4704. doi: 10.1073/pnas.051628998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Currie KP. Inhibition of Ca2+ Channels and adrenal catecholamine release by G protein coupled receptors. Cell Mol Neurobiol. 2010 doi: 10.1007/s10571-010-9596-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Currie KP, Fox AP. ATP serves as a negative feedback inhibitor of voltage-gated Ca2+ channel currents in cultured bovine adrenal chromaffin cells. Neuron. 1996;16:1027–1036. doi: 10.1016/s0896-6273(00)80126-9. [DOI] [PubMed] [Google Scholar]
  • 58.Womack MD, McCleskey EW. Interaction of opioids and membrane potential to modulate Ca2+ channels in rat dorsal root ganglion neurons. J Neurophysiol. 1995;73:1793–1798. doi: 10.1152/jn.1995.73.5.1793. [DOI] [PubMed] [Google Scholar]
  • 59.Williams S, Serafin M, Muhlethaler M, Bernheim L. Facilitation of N-type calcium current is dependent on the frequency of action potential-like depolarizations in dissociated cholinergic basal forebrain neurons of the guinea pig. J Neurosci. 1997;17:1625–1632. doi: 10.1523/JNEUROSCI.17-05-01625.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Park D, Dunlap K. Dynamic regulation of calcium influx by G-proteins, action potential waveform and neuronal firing frequency. J Neurosci. 1998;18:6757–6766. doi: 10.1523/JNEUROSCI.18-17-06757.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Tosetti P, Taglietti V, Toselli M. Action-potential-like depolarizations relieve opioid inhibition of N-type Ca2+ channels in NG108-15 cells. Pflugers Arch. 1999;437:441–448. doi: 10.1007/s004240050799. [DOI] [PubMed] [Google Scholar]
  • 62.Brody DL, Patil PG, Mulle JG, Snutch TP, Yue DT. Bursts of action potential waveforms relieve G-protein inhibition of recombinant P/Q-type Ca2+ channels in HEK 293 cells. J Physiol. 1997;499:637–644. doi: 10.1113/jphysiol.1997.sp021956. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Currie KP, Fox AP. Differential facilitation of N- and P/Q-type calcium channels during trains of action potential-like waveforms. J Physiol. 2002;539:419–431. doi: 10.1113/jphysiol.2001.013206. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Brody DL, Yue DT. Relief of G-protein inhibition of calcium channels and short-term synaptic facilitation in cultured hippocampal neurons. J Neurosci. 2000;20:889–898. doi: 10.1523/JNEUROSCI.20-03-00889.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Golard A, Siegelbaum SA. Kinetic basis for the voltage-dependent inhibition of N-type calcium current by somatostatin and norepinephrine in chick sympathetic neurons. J Neurosci. 1993;13:3884–3894. doi: 10.1523/JNEUROSCI.13-09-03884.1993. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Elmslie KS, Jones SW. Concentration dependence of neurotransmitter effects on calcium current kinetics in frog sympathetic neurones. J Physiol. 1994;481:35–46. doi: 10.1113/jphysiol.1994.sp020417. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Delmas P, Brown DA, Dayrell M, Abogadie FC, Caulfield MP, Buckley NJ. On the role of endogenous G-protein β γ subunits in N-type Ca2+ current inhibition by neurotransmitters in rat sympathetic neurones. J Physiol. 1998;506:319–329. doi: 10.1111/j.1469-7793.1998.319bw.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Zamponi GW, Snutch TP. Modulation of voltage-dependent calcium channels by G proteins. Curr Opin Neurobiol. 1998;8:351–356. doi: 10.1016/s0959-4388(98)80060-3. [DOI] [PubMed] [Google Scholar]
  • 69.Forscher P, Oxford GS, Schulz D. Noradrenaline modulates calcium channels in avian dorsal root ganglion cells through tight receptor-channel coupling. J Physiol. 1986;379:131–144. doi: 10.1113/jphysiol.1986.sp016244. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Bernheim L, Beech DJ, Hille B. A diffusible second messenger mediates one of the pathways coupling receptors to calcium channels in rat sympathetic neurons. Neuron. 1991;6:859–867. doi: 10.1016/0896-6273(91)90226-p. [DOI] [PubMed] [Google Scholar]
  • 71.Patil PG, de Leon M, Reed RR, Dubel S, Snutch TP, Yue DT. Elementary events underlying voltage-dependent G-protein inhibition of N-type calcium channels. Biophys J. 1996;71:2509–2521. doi: 10.1016/S0006-3495(96)79444-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Boland LM, Bean BP. Modulation of N-type calcium channels in bullfrog sympathetic neurons by luteinizing hormone-releasing hormone: kinetics and voltage dependence. J Neurosci. 1993;13:516–533. doi: 10.1523/JNEUROSCI.13-02-00516.1993. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Jones SW, Elmslie KS. Transmitter modulation of neuronal calcium channels. J Membr Biol. 1997;155:1–10. doi: 10.1007/s002329900153. [DOI] [PubMed] [Google Scholar]
  • 74.Rebolledo-Antunez S, Farias JM, Arenas I, Garcia DE. Gating charges per channel of CaV2.2 channels are modified by G protein activation in rat sympathetic neurons. Arch Biochem Biophys. 2009;486:51–57. doi: 10.1016/j.abb.2009.04.002. [DOI] [PubMed] [Google Scholar]
  • 75.Hernandez-Ochoa EO, Garcia-Ferreiro RE, Garcia DE. G protein activation inhibits gating charge movement in rat sympathetic neurons. American journal of physiology. 2007;292:2226–2238. doi: 10.1152/ajpcell.00540.2006. [DOI] [PubMed] [Google Scholar]
  • 76.McDavid S, Currie KP. G-proteins modulate cumulative inactivation of N-type (CaV2.2) calcium channels. J Neurosci. 2006;26:13373–13383. doi: 10.1523/JNEUROSCI.3332-06.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Weiss N, Tadmouri A, Mikati M, Ronjat M, De Waard M. Importance of voltage-dependent inactivation in N-type calcium channel regulation by G-proteins. Pflugers Arch. 2007;454:115–129. doi: 10.1007/s00424-006-0184-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Cens T, Rousset M, Leyris JP, Fesquet P, Charnet P. Voltage- and calcium-dependent inactivation in high voltage-gated Ca2+ channels. Prog Biophys Mol Biol. 2006;90:104–117. doi: 10.1016/j.pbiomolbio.2005.05.013. [DOI] [PubMed] [Google Scholar]
  • 79.Stotz SC, Zamponi GW. Structural determinants of fast inactivation of high voltage-activated Ca2+ channels. Trends Neurosci. 2001;24:176–181. doi: 10.1016/s0166-2236(00)01738-0. [DOI] [PubMed] [Google Scholar]
  • 80.Hering S, Berjukow S, Sokolov S, Marksteiner R, Weiss RG, Kraus R, et al. Molecular determinants of inactivation in voltage-gated Ca2+ channels. J Physiol. 2000;528:237–249. doi: 10.1111/j.1469-7793.2000.t01-1-00237.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81.Tadross MR, Ben Johny M, Yue DT. Molecular endpoints of Ca2+/calmodulin- and voltage-dependent inactivation of CaV1.3 channels. J Gen Physiol. 2010;135:197–215. doi: 10.1085/jgp.200910308. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Findeisen F, Minor DL., Jr Disruption of the IS6-AID linker affects voltage-gated calcium channel inactivation and facilitation. J Gen Physiol. 2009;133:327–343. doi: 10.1085/jgp.200810143. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Herlitze S, Hockerman GH, Scheuer T, Catterall WA. Molecular determinants of inactivation and G protein modulation in the intracellular loop connecting domains I and II of the calcium channel α1A subunit. Proc Natl Acad Sci USA. 1997;94:1512–1516. doi: 10.1073/pnas.94.4.1512. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.De Waard M, Liu H, Walker D, Scott VE, Gurnett CA, Campbell KP. Direct binding of G-protein βγ complex to voltage-dependent calcium channels. Nature. 1997;385:446–450. doi: 10.1038/385446a0. [DOI] [PubMed] [Google Scholar]
  • 85.De Waard M, Hering J, Weiss N, Feltz A. How do G proteins directly control neuronal Ca2+ channel function? Trends Pharmacol Sci. 2005;26:427–436. doi: 10.1016/j.tips.2005.06.008. [DOI] [PubMed] [Google Scholar]
  • 86.Schiff ML, Siderovski DP, Jordan JD, Brothers G, Snow B, De Vries L, et al. Tyrosine-kinase-dependent recruitment of RGS12 to the N-type calcium channel. Nature. 2000;408:723–727. doi: 10.1038/35047093. [DOI] [PubMed] [Google Scholar]
  • 87.Peterson BZ, DeMaria CD, Adelman JP, Yue DT. Calmodulin is the Ca2+ sensor for Ca2+-dependent inactivation of L-type calcium channels. Neuron. 1999;22:549–558. doi: 10.1016/s0896-6273(00)80709-6. [DOI] [PubMed] [Google Scholar]
  • 88.Zuhlke RD, Pitt GS, Deisseroth K, Tsien RW, Reuter H. Calmodulin supports both inactivation and facilitation of L-type calcium channels. Nature. 1999;399:159–162. doi: 10.1038/20200. [DOI] [PubMed] [Google Scholar]
  • 89.Liang H, DeMaria CD, Erickson MG, Mori MX, Alseikhan BA, Yue DT. Unified mechanisms of Ca2+ regulation across the Ca2+ channel family. Neuron. 2003;39:951–960. doi: 10.1016/s0896-6273(03)00560-9. [DOI] [PubMed] [Google Scholar]
  • 90.Lee A, Wong ST, Gallagher D, Li B, Storm DR, Scheuer T, et al. Ca2+/calmodulin binds to and modulates P/Qtype calcium channels. Nature. 1999;399:155–159. doi: 10.1038/20194. [DOI] [PubMed] [Google Scholar]
  • 91.Lee A, Zhou H, Scheuer T, Catterall WA. Molecular determinants of Ca2+/calmodulin-dependent regulation of CaV2.1 channels. Proc Natl Acad Sci USA. 2003;100:16059–16064. doi: 10.1073/pnas.2237000100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Dick IE, Tadross MR, Liang H, Tay LH, Yang W, Yue DT. A modular switch for spatial Ca2+ selectivity in the calmodulin regulation of CaV channels. Nature. 2008;451:830–834. doi: 10.1038/nature06529. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.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. doi: 10.1074/jbc.M005881200. [DOI] [PubMed] [Google Scholar]
  • 94.Benmocha A, Almagor L, Oz S, Hirsch JA, Dascal N. Characterization of the calmodulin-binding site in the N terminus of CaV1.2. Channels (Austin) 2009;3:337–342. doi: 10.4161/chan.3.5.9686. [DOI] [PubMed] [Google Scholar]
  • 95.Liu M, Yu B, Nakanishi O, Wieland T, Simon M. The Ca2+-dependent binding of calmodulin to an N-terminal motif of the heterotrimeric G protein β subunit. J Biol Chem. 1997;272:18801–18807. doi: 10.1074/jbc.272.30.18801. [DOI] [PubMed] [Google Scholar]
  • 96.Toth PT, Shekter LR, Ma GH, Philipson LH, Miller RJ. Selective G-protein regulation of neuronal calcium channels. J Neurosci. 1996;16:4617–4624. doi: 10.1523/JNEUROSCI.16-15-04617.1996. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Page KM, Canti C, Stephens GJ, Berrow NS, Dolphin AC. Identification of the amino terminus of neuronal Ca2+ channel α1 subunits α1B and α1E as an essential determinant of G-protein modulation. J Neurosci. 1998;18:4815–4824. doi: 10.1523/JNEUROSCI.18-13-04815.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Meza U, Adams B. G-Protein-dependent facilitation of neuronal α1A, α1B and α1E Ca channels. J Neurosci. 1998;18:5240–5252. doi: 10.1523/JNEUROSCI.18-14-05240.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Bourinet E, Soong TW, Stea A, Snutch TP. Determinants of the G protein-dependent opioid modulation of neuronal calcium channels. Proc Natl Acad Sci USA. 1996;93:1486–1491. doi: 10.1073/pnas.93.4.1486. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Zhang JF, Ellinor PT, Aldrich RW, Tsien RW. Multiple structural elements in voltage-dependent Ca2+ channels support their inhibition by G proteins. Neuron. 1996;17:991–1003. doi: 10.1016/s0896-6273(00)80229-9. [DOI] [PubMed] [Google Scholar]
  • 101.Currie KPM, Fox AP. Comparison of N- and P/Q-type voltage-gated calcium channel current inhibition. J Neurosci. 1997;17:4570–4579. doi: 10.1523/JNEUROSCI.17-12-04570.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102.Currie KPM, Fox AP. Differential facilitation of N- and P/Q-type calcium channels during trains of action potential-like waveforms. J Physiol. 2002;539:419–431. doi: 10.1113/jphysiol.2001.013206. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 103.Arnot MI, Stotz SC, Jarvis SE, Zamponi GW. Differential modulation of N-type 1B and P/Q-type 1A calcium channels by different G protein subunit isoforms. J Physiol. 2000;527:203–212. doi: 10.1111/j.1469-7793.2000.00203.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104.Agler HL, Evans J, Colecraft HM, Yue DT. Custom distinctions in the interaction of G-protein β subunits with N-type (CaV2.2) versus P/Q-type (CaV2.1) calcium channels. J Gen Physiol. 2003;121:495–510. doi: 10.1085/jgp.200208770. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105.Swartz KJ. Modulation of Ca2+ channels by protein kinase C in rat central and peripheral neurons: disruption of G protein-mediated inhibition. Neuron. 1993;11:305–320. doi: 10.1016/0896-6273(93)90186-u. [DOI] [PubMed] [Google Scholar]
  • 106.Barrett CF, Rittenhouse AR. Modulation of N-type calcium channel activity by G-proteins and protein kinase C. J Gen Physiol. 2000;115:277–286. doi: 10.1085/jgp.115.3.277. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.Simen AA, Lee CC, Simen BB, Bindokas VP, Miller RJ. The C terminus of the Ca channel α1B subunit mediates selective inhibition by G-protein-coupled receptors. J Neurosci. 2001;21:7587–7597. doi: 10.1523/JNEUROSCI.21-19-07587.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108.Bertaso F, Ward RJ, Viard P, Milligan G, Dolphin AC. Mechanism of action of Gq to inhibit Gβ γ modulation of CaV2.2 calcium channels: probed by the use of receptor-Gα tandems. Mol Pharmacol. 2003;63:832–843. doi: 10.1124/mol.63.4.832. [DOI] [PubMed] [Google Scholar]
  • 109.Zamponi GW, Bourinet E, Nelson D, Nargeot J, Snutch TP. Crosstalk between G proteins and protein kinase C mediated by the calcium channel α1 subunit. Nature. 1997;385:442–446. doi: 10.1038/385442a0. [DOI] [PubMed] [Google Scholar]
  • 110.Hamid J, Nelson D, Spaetgens R, Dubel SJ, Snutch TP, Zamponi GW. Identification of an integration center for cross-talk between protein kinase C and G protein modulation of N-type calcium channels. J Biol Chem. 1999;274:6195–6202. doi: 10.1074/jbc.274.10.6195. [DOI] [PubMed] [Google Scholar]
  • 111.Cooper CB, Arnot MI, Feng ZP, Jarvis SE, Hamid J, Zamponi GW. Cross-talk between G-protein and protein kinase C modulation of N-type calcium channels is dependent on the G-protein b subunit isoform. J Biol Chem. 2000;275:40777–40781. doi: 10.1074/jbc.C000673200. [DOI] [PubMed] [Google Scholar]
  • 112.Wu X, Kushwaha N, Albert PR, Penington NJ. A critical protein kinase C phosphorylation site on the 5-HT(1A) receptor controlling coupling to N-type calcium channels. J Physiol. 2002;538:41–51. doi: 10.1113/jphysiol.2001.012668. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113.Qin N, Platano D, Olcese R, Stefani E, Birnbaumer L. Direct interaction of gβγ with a C-terminal gβγ-binding domain of the Ca2+ channel α1 subunit is responsible for channel inhibition by G protein-coupled receptors. Proc Natl Acad Sci USA. 1997;94:8866–8871. doi: 10.1073/pnas.94.16.8866. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Li B, Zhong H, Scheuer T, Catterall WA. Functional role of a C-terminal Gβγ-binding domain of CaV2.2 channels. Mol Pharmacol. 2004;66:761–769. doi: 10.1124/mol.66.3.. [DOI] [PubMed] [Google Scholar]
  • 115.Furukawa T, Nukada T, Mori Y, Wakamori M, Fujita Y, Ishida H, et al. Differential interactions of the C terminus and the cytoplasmic I–II loop of neuronal Ca2+ channels with G-protein α and βγ subunits. I. Molecular determination. J Biol Chem. 1998;273:17585–17594. doi: 10.1074/jbc.273.28.17585. [DOI] [PubMed] [Google Scholar]
  • 116.Evans RM, Zamponi GW. Presynaptic Ca2+ channels—integration centers for neuronal signaling pathways. Trends Neurosci. 2006;29:617–624. doi: 10.1016/j.tins.2006.08.006. [DOI] [PubMed] [Google Scholar]
  • 117.Tedford HW, Kisilevsky AE, Vieira LB, Varela D, Chen L, Zamponi GW. Scanning mutagenesis of the I–II loop of the CaV2.2 calcium channel identifies residues Arginine 376 and Valine 416 as molecular determinants of voltage dependent G protein inhibition. Mol Brain. 2010;3:6. doi: 10.1186/1756-6606-3-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 118.Pragnell M, De Waard M, Mori Y, Tanabe T, Snutch TP, Campbell KP. Calcium channel β-subunit binds to a conserved motif in the I–II cytoplasmic linker of the α1-subunit. Nature. 1994;368:67–70. doi: 10.1038/368067a0. [DOI] [PubMed] [Google Scholar]
  • 119.Bell DC, Butcher AJ, Berrow NS, Page KM, Brust PF, Nesterova A, et al. Biophysical properties, pharmacology and modulation of human, neuronal L-type (α1D, CaV1.3) voltage-dependent calcium currents. J Neurophysiol. 2001;85:816–827. doi: 10.1152/jn.2001.85.2.816. [DOI] [PubMed] [Google Scholar]
  • 120.Zhang Y, Chen YH, Bangaru SD, He L, Abele K, Tanabe S, et al. Origin of the voltage dependence of G-protein regulation of P/Q-type Ca2+ channels. J Neurosci. 2008;28:14176–14188. doi: 10.1523/JNEUROSCI.1350-08.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 121.Canti C, Page KM, Stephens GJ, Dolphin AC. Identification of residues in the N terminus of α1B critical for inhibition of the voltage-dependent calcium channel by Gβ γ. J Neurosci. 1999;19:6855–6864. doi: 10.1523/JNEUROSCI.19-16-06855.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Agler HL, Evans J, Tay LH, Anderson MJ, Colecraft HM, Yue DT. G protein-gated inhibitory module of N-type (CaV2.2) Ca2+ channels. Neuron. 2005;46:891–904. doi: 10.1016/j.neuron.2005.05.011. [DOI] [PubMed] [Google Scholar]
  • 123.Leroy J, Richards MW, Butcher AJ, Nieto-Rostro M, Pratt WS, Davies A, et al. Interaction via a key tryptophan in the I–II linker of N-type calcium channels is required for β1 but not for palmitoylated β2, implicating an additional binding site in the regulation of channel voltage-dependent properties. J Neurosci. 2005;25:6984–6996. doi: 10.1523/JNEUROSCI.1137-05.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124.Dresviannikov AV, Page KM, Leroy J, Pratt WS, Dolphin AC. Determinants of the voltage dependence of G protein modulation within calcium channel β subunits. Pflugers Arch. 2009;457:743–756. doi: 10.1007/s00424-008-0549-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 125.Stephens GJ, Canti C, Page KM, Dolphin AC. Role of domain I of neuronal Ca2+ channel α1 subunits in G protein modulation. J Physiol. 1998;509:163–169. doi: 10.1111/j.1469-7793.1998.163bo.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 126.Simen AA, Miller RJ. Structural features determining differential receptor regulation of neuronal Ca channels. J Neurosci. 1998;18:3689–3698. doi: 10.1523/JNEUROSCI.18-10-03689.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 127.Page KM, Heblich F, Margas W, Pratt WS, Nieto-Rostro M, Chaggar K, et al. N terminus is key to the dominant negative suppression of CaV2 calcium channels: implications for episodic ataxia type 2. J Biol Chem. 2010;285:835–844. doi: 10.1074/jbc.M109.065045. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 128.Wall MA, Coleman DE, Lee E, Iniguez-Lluhi JA, Posner BA, Gilman AG, et al. The structure of the G protein heterotrimer Giα1β1γ2. Cell. 1995;83:1047–1058. doi: 10.1016/0092-8674(95)90220-1. [DOI] [PubMed] [Google Scholar]
  • 129.Sondek J, Bohm A, Lambright DG, Hamm HE, Sigler PB. Crystal structure of a G-protein βγ dimer at 2.1 A resolution. Nature. 1996;379:369–374. doi: 10.1038/379369a0. [DOI] [PubMed] [Google Scholar]
  • 130.Lambright DG, Sondek J, Bohm A, Skiba NP, Hamm HE, Sigler PB. The 2.0 A crystal structure of a heterotrimeric G protein. Nature. 1996;379:311–319. doi: 10.1038/379311a0. [DOI] [PubMed] [Google Scholar]
  • 131.Gaudet R, Bohm A, Sigler PB. Crystal structure at 2.4 angstroms resolution of the complex of transducin βγ and its regulator, phosducin. Cell. 1996;87:577–588. doi: 10.1016/s0092-8674(00)81376-8. [DOI] [PubMed] [Google Scholar]
  • 132.Lodowski DT, Pitcher JA, Capel WD, Lefkowitz RJ, Tesmer JJ. Keeping G proteins at bay: a complex between G protein-coupled receptor kinase 2 and Gβγ. Science. 2003;300:1256–1262. doi: 10.1126/science.1082348. [DOI] [PubMed] [Google Scholar]
  • 133.Smrcka AV. G protein βγ subunits: central mediators of G protein-coupled receptor signaling. Cell Mol Life Sci. 2008;65:2191–2214. doi: 10.1007/s00018-008-8006-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 134.Ford CE, Skiba NP, Bae H, Daaka Y, Reuveny E, Shekter LR, et al. Molecular basis for interactions of G protein βγ subunits with effectors. Science. 1998;280:1271–1274. doi: 10.1126/science.280.5367.1271. [DOI] [PubMed] [Google Scholar]
  • 135.Tedford HW, Kisilevsky AE, Peloquin JB, Zamponi GW. Scanning mutagenesis reveals a role for serine 189 of the heterotrimeric G-protein β1 subunit in the inhibition of N-type calcium channels. J Neurophysiol. 2006;96:465–470. doi: 10.1152/jn.00216.2006. [DOI] [PubMed] [Google Scholar]
  • 136.Doering CJ, Kisilevsky AE, Feng ZP, Arnot MI, Peloquin J, Hamid J, et al. A single Gβ subunit locus controls crosstalk between PKC and G protein regulation of N-type calcium channels. J Biol Chem. 2004 doi: 10.1074/jbc.M308693200. [DOI] [PubMed] [Google Scholar]
  • 137.Mirshahi T, Mittal V, Zhang H, Linder ME, Logothetis DE. Distinct sites on G protein βγ subunits regulate different effector functions. J Biol Chem. 2002;277:36345–36350. doi: 10.1074/jbc.M205359200. [DOI] [PubMed] [Google Scholar]
  • 138.Li X, Hummer A, Han J, Xie M, Melnik-Martinez K, Moreno RL, et al. G protein β2 subunit-derived peptides for inhibition and induction of G protein pathways. Examination of voltage-gated Ca2+ and G protein inwardly rectifying K+ channels. J Biol Chem. 2005;280:23945–23959. doi: 10.1074/jbc.M414078200. [DOI] [PubMed] [Google Scholar]
  • 139.Zhou JY, Siderovski DP, Miller RJ. Selective regulation of N-type Ca channels by different combinations of G-protein β/γ subunits and RGS proteins. J Neurosci. 2000;20:7143–7148. doi: 10.1523/JNEUROSCI.20-19-07143.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 140.Blake BL, Wing MR, Zhou JY, Lei Q, Hillmann JR, Behe CI, et al. Gβ association and effector interaction selectivities of the divergent Gγ subunit Gγ(13) J Biol Chem. 2001;276:49267–48274. doi: 10.1074/jbc.M106565200. [DOI] [PubMed] [Google Scholar]
  • 141.Fathallah M, Sandoz G, Mabrouk K, Geib S, Urbani J, Villaz M, et al. Modelling of the III–IV loop, a domain involved in calcium channel CaV2.1 inactivation, highlights a structural homology with the γ subunit of G proteins. Eur J Neurosci. 2002;16:219–228. doi: 10.1046/j.1460-9568.2002.02074.x. [DOI] [PubMed] [Google Scholar]
  • 142.Bonacci TM, Mathews JL, Yuan C, Lehmann DM, Malik S, Wu D, et al. Differential targeting of Gβγ-subunit signaling with small molecules. Science. 2006;312:443–446. doi: 10.1126/science.1120378. [DOI] [PubMed] [Google Scholar]
  • 143.Mathews JL, Smrcka AV, Bidlack JM. A novel Gβγ-subunit inhibitor selectively modulates mu-opioiddependent antinociception and attenuates acute morphine-induced antinociceptive tolerance and dependence. J Neurosci. 2008;28:12183–12189. doi: 10.1523/JNEUROSCI.2326-08.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 144.Casey LM, Pistner AR, Belmonte SL, Migdalovich D, Stolpnik O, Nwakanma FE, et al. Small molecule disruption of Gβγ signaling inhibits the progression of heart failure. Circ Res. 2010;107:532–539. doi: 10.1161/CIRCRESAHA.110.217075. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 145.Meir A, Bell DC, Stephens GJ, Page KM, Dolphin AC. Calcium channel β subunit promotes voltage-dependent modulation of a1B by Gβγ. Biophys J. 2000;79:731–746. doi: 10.1016/S0006-3495(00)76331-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 146.Sandoz G, Lopez-Gonzalez I, Stamboulian S, Weiss N, Arnoult C, De Waard M. Repositioning of charged I–II loop amino acid residues within the electric field by β subunit as a novel working hypothesis for the control of fast P/Q calcium channel inactivation. Eur J Neurosci. 2004;19:1759–1772. doi: 10.1111/j.1460-9568.2004.03216.x. [DOI] [PubMed] [Google Scholar]
  • 147.Hummer A, Delzeith O, Gomez SR, Moreno RL, Mark MD, Herlitze S. Competitive and synergistic interactions of G protein β2 and Ca2+ channel β1b subunits with CaV2.1 channels, revealed by mammalian two-hybrid and fluorescence resonance energy transfer measurements. J Biol Chem. 2003;278:49386–49400. doi: 10.1074/jbc.M306645200. [DOI] [PubMed] [Google Scholar]
  • 148.Canti C, Bogdanov Y, Dolphin AC. Interaction between G proteins and accessory subunits in the regulation of 1B calcium channels in Xenopus oocytes. J Physiol. 2000;527:419–432. doi: 10.1111/j.1469-7793.2000.t01-1-00419.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 149.Feng ZP, Arnot MI, Doering CJ, Zamponi GW. Calcium channel β subunits differentially regulate the inhibition of N-type channels by individual Gβ isoforms. J Biol Chem. 2001;276:45051–45058. doi: 10.1074/jbc.M107784200. [DOI] [PubMed] [Google Scholar]
  • 150.Richards MW, Leroy J, Pratt WS, Dolphin AC. The HOOK-domain between the SH3 and the GK domains of Cavβ subunits contains key determinants controlling calcium channel inactivation. Channels (Austin) 2007;1:92–101. doi: 10.4161/chan.4145. [DOI] [PubMed] [Google Scholar]
  • 151.Vitko I, Shcheglovitov A, Baumgart JP, Arias O, II, Murbartian J, Arias JM, et al. Orientation of the calcium channel β relative to the α12.2 subunit is critical for its regulation of channel activity. PLoS One. 2008;3:3560. doi: 10.1371/journal.pone.0003560. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 152.Heneghan JF, Mitra-Ganguli T, Stanish LF, Liu L, Zhao R, Rittenhouse AR. The Ca2+ channel β subunit determines whether stimulation of Gq-coupled receptors enhances or inhibits N current. J Gen Physiol. 2009;134:369–384. doi: 10.1085/jgp.200910203. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 153.Mitra-Ganguli T, Vitko I, Perez-Reyes E, Rittenhouse AR. Orientation of palmitoylated CaVβ2a relative to CaV2.2 is critical for slow pathway modulation of N-type Ca2+ current by tachykinin receptor activation. J Gen Physiol. 2009;134:385–396. doi: 10.1085/jgp.200910204. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 154.Arias JM, Murbartian J, Vitko I, Lee JH, Perez-Reyes E. Transfer of β subunit regulation from high to low voltage-gated Ca2+ channels. FEBS Lett. 2005;579:3907–3912. doi: 10.1016/j.febslet.2005.06.008. [DOI] [PubMed] [Google Scholar]
  • 155.Diaz E. Regulation of AMPA receptors by transmembrane accessory proteins. Eur J Neurosci. 2010;32:261–268. doi: 10.1111/j.1460-9568.2010.07357.x. [DOI] [PubMed] [Google Scholar]
  • 156.Tselnicker I, Tsemakhovich VA, Dessauer CW, Dascal N. Stargazin modulates neuronal voltage-dependent Ca2+ channel CaV2.2 by a Gβγ-dependent mechanism. J Biol Chem. 2010;285:20462–20471. doi: 10.1074/jbc.M110.121277. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 157.Davies JN, Zamponi GW. Old proteins, developing roles: The regulation of calcium channels by synaptic proteins. Channels (Austin) 2008;2:130–138. doi: 10.4161/chan.2.2.6214. [DOI] [PubMed] [Google Scholar]
  • 158.Jarvis SE, Zamponi GW. Masters or slaves? Vesicle release machinery and the regulation of presynaptic calcium channels. Cell Calcium. 2005;37:483–488. doi: 10.1016/j.ceca.2005.01.017. [DOI] [PubMed] [Google Scholar]
  • 159.Sheng ZH, Rettig J, Takahashi M, Catterall WA. Identification of a syntaxin-binding site on N-type calcium channels. Neuron. 1994;13:1303–1313. doi: 10.1016/0896-6273(94)90417-0. [DOI] [PubMed] [Google Scholar]
  • 160.Bezprozvanny I, Scheller RH, Tsien RW. Functional impact of syntaxin on gating of N-type and Q-type calcium channels. Nature. 1995;378:623–626. doi: 10.1038/378623a0. [DOI] [PubMed] [Google Scholar]
  • 161.Sheng ZH, Yokoyama CT, Catterall WA. Interaction of the synprint site of N-type Ca2+ channels with the C2B domain of synaptotagmin I. Proc Natl Acad Sci USA. 1997;94:5405–5410. doi: 10.1073/pnas.94.10.5405. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 162.Jarvis SE, Barr W, Feng ZP, Hamid J, Zamponi GW. Molecular determinants of syntaxin 1 modulation of N-type calcium channels. J Biol Chem. 2002;277:44399–44407. doi: 10.1074/jbc.M206902200. [DOI] [PubMed] [Google Scholar]
  • 163.Wiser O, Bennett MK, Atlas D. Functional interaction of syntaxin and SNAP-25 with voltage-sensitive L- and N-type Ca2+ channels. EMBO J. 1996;15:4100–4110. [PMC free article] [PubMed] [Google Scholar]
  • 164.Yokoyama CT, Myers SJ, Fu J, Mockus SM, Scheuer T, Catterall WA. Mechanism of SNARE protein binding and regulation of CaV2 channels by phosphorylation of the synaptic protein interaction site. Mol Cell Neurosci. 2005;28:1–17. doi: 10.1016/j.mcn.2004.08.019. [DOI] [PubMed] [Google Scholar]
  • 165.Jarvis SE, Magga JM, Beedle AM, Braun JE, Zamponi GW. G protein modulation of N-type calcium channels is facilitated by physical interactions between syntaxin 1A and G βγ. J Biol Chem. 2000;275:6388–6394. doi: 10.1074/jbc.275.9.6388. [DOI] [PubMed] [Google Scholar]
  • 166.Jarvis SE, Zamponi GW. Interactions between pre-synaptic Ca2+ channels, cytoplasmic messengers and proteins of the synaptic vesicle release complex. Trends Pharmacol Sci. 2001;22:519–525. doi: 10.1016/s0165-6147(00)01800-9. [DOI] [PubMed] [Google Scholar]
  • 167.Lu Q, AtKisson MS, Jarvis SE, Feng ZP, Zamponi GW, Dunlap K. Syntaxin 1A supports voltage-dependent inhibition of α1B Ca2+ channels by Gbg in chick sensory neurons. J Neurosci. 2001;21:2949–2957. doi: 10.1523/JNEUROSCI.21-09-02949.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 168.Stanley EF, Mirotznik RR. Cleavage of syntaxin prevents G-protein regulation of presynaptic calcium channels. Nature. 1997;385:340–343. doi: 10.1038/385340a0. [DOI] [PubMed] [Google Scholar]
  • 169.Silinsky EM. Modulation of calcium currents is eliminated after cleavage of a strategic component of the mammalian secretory apparatus. J Physiol. 2005;566:681–688. doi: 10.1113/jphysiol.2005.090647. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 170.Magga JM, Jarvis SE, Arnot MI, Zamponi GW, Braun JE. Cysteine string protein regulates G protein modulation of N-type calcium channels. Neuron. 2000;28:195–204. doi: 10.1016/s0896-6273(00)00096-9. [DOI] [PubMed] [Google Scholar]
  • 171.Miller LC, Swayne LA, Kay JG, Feng ZP, Jarvis SE, Zamponi GW, et al. Molecular determinants of cysteine string protein modulation of N-type calcium channels. J Cell Sci. 2003;116:2967–2974. doi: 10.1242/jcs.00595. [DOI] [PubMed] [Google Scholar]
  • 172.Sethakorn N, Yau DM, Dulin NO. Non-canonical functions of RGS proteins. Cell Signal. 2010;22:1274–1281. doi: 10.1016/j.cellsig.2010.03.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 173.Richman RW, Strock J, Hains MD, Cabanilla NJ, Lau KK, Siderovski DP, et al. RGS12 interacts with the SNARE-binding region of the Cav2.2 calcium channel. J Biol Chem. 2005;280:1521–1528. doi: 10.1074/jbc.M406607200. [DOI] [PubMed] [Google Scholar]
  • 174.Tosetti P, Pathak N, Jacob MH, Dunlap K. RGS3 mediates a calcium-dependent termination of G protein signaling in sensory neurons. Proc Natl Acad Sci USA. 2003;100:7337–7342. doi: 10.1073/pnas.1231837100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 175.Kammermeier PJ, Ikeda SR. Expression of RGS2 alters the coupling of metabotropic glutamate receptor 1a to M-type K+ and N-type Ca2+ channels. Neuron. 1999;22:819–829. doi: 10.1016/s0896-6273(00)80740-0. [DOI] [PubMed] [Google Scholar]
  • 176.Han J, Mark MD, Li X, Xie M, Waka S, Rettig J, et al. RGS2 determines short-term synaptic plasticity in hippocampal neurons by regulating Gi/o-mediated inhibition of presynaptic Ca2+ channels. Neuron. 2006;51:575–586. doi: 10.1016/j.neuron.2006.07.012. [DOI] [PubMed] [Google Scholar]
  • 177.Ding J, Guzman JN, Tkatch T, Chen S, Goldberg JA, Ebert PJ, et al. RGS4-dependent attenuation of M4 autoreceptor function in striatal cholinergic interneurons following dopamine depletion. Nat Neurosci. 2006;9:832–842. doi: 10.1038/nn1700. [DOI] [PubMed] [Google Scholar]
  • 178.Sjogren B, Blazer LL, Neubig RR. Regulators of G protein signaling proteins as targets for drug discovery. Prog Mol Biol Transl Sci. 2010;91:81–119. doi: 10.1016/S1877-1173(10)91004-1. [DOI] [PubMed] [Google Scholar]
  • 179.Thapliyal A, Bannister RA, Hanks C, Adams BA. The monomeric G proteins AGS1 and Rhes selectively influence Gαi-dependent signaling to modulate N-type (CaV2.2) calcium channels. Am J Physiol Cell Physiol. 2008;295:1417–1426. doi: 10.1152/ajpcell.00341.2008. [DOI] [PubMed] [Google Scholar]
  • 180.Kammermeier PJ, Xiao B, Tu JC, Worley PF, Ikeda SR. Homer proteins regulate coupling of group I metabotropic glutamate receptors to N-type calcium and M-type potassium channels. J Neurosci. 2000;20:7238–7245. doi: 10.1523/JNEUROSCI.20-19-07238.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 181.Filippov AK, Simon J, Barnard EA, Brown DA. The scaffold protein NHERF2 determines the coupling of P2Y1 nucleotide and mGluR5 glutamate receptor to different ion channels in neurons. J Neurosci. 2010;30:11068–11072. doi: 10.1523/JNEUROSCI.2597-10.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 182.Beedle AM, McRory JE, Poirot O, Doering CJ, Altier C, Barrere C, et al. Agonist-independent modulation of N-type calcium channels by ORL1 receptors. Nat Neurosci. 2004;7:118–125. doi: 10.1038/nn1180. [DOI] [PubMed] [Google Scholar]
  • 183.Evans RM, You H, Hameed S, Altier C, Mezghrani A, Bourinet E, et al. Heterodimerization of ORL1 and opioid receptors and its consequences for N-type calcium channel regulation. J Biol Chem. 2010;285:1032–1040. doi: 10.1074/jbc.M109.040634. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 184.Kisilevsky AE, Zamponi GW. D2 dopamine receptors interact directly with N-type calcium channels and regulate channel surface expression levels. Channels (Austin) 2008;2:269–277. doi: 10.4161/chan.2.4.6402. [DOI] [PubMed] [Google Scholar]
  • 185.Pietrobon D. CaV2.1 channelopathies. Pflugers Arch. 2010;460:375–393. doi: 10.1007/s00424-010-0802-8. [DOI] [PubMed] [Google Scholar]
  • 186.Lorenzon NM, Beam KG. Disease causing mutations of calcium channels. Channels (Austin) 2008;2:163–179. doi: 10.4161/chan.2.3.5950. [DOI] [PubMed] [Google Scholar]
  • 187.Melliti K, Grabner M, Seabrook GR. The familial hemiplegic migraine mutation R192Q reduces G-protein-mediated inhibition of P/Q-type (CaV2.1) calcium channels expressed in human embryonic kidney cells. J Physiol. 2003;546:337–347. doi: 10.1113/jphysiol.2002.026716. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 188.Serra SA, Fernandez-Castillo N, Macaya A, Cormand B, Valverde MA, Fernandez-Fernandez JM. The hemiplegic migraine-associated Y1245C mutation in CACNA1A results in a gain of channel function due to its effect on the voltage sensor and G-protein-mediated inhibition. Pflugers Arch. 2009;458:489–502. doi: 10.1007/s00424-009-0637-3. [DOI] [PubMed] [Google Scholar]
  • 189.Weiss N, Sandoval A, Felix R, Van den Maagdenberg A, De Waard M. The S218L familial hemiplegic migraine mutation promotes deinhibition of CaV2.1 calcium channels during direct G-protein regulation. Pflugers Arch. 2008;457:315–326. doi: 10.1007/s00424-008-0541-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 190.Kammermeier PJ, Ruiz-Velasco V, Ikeda SR. A voltage-independent calcium current inhibitory pathway activated by muscarinic agonists in rat sympathetic neurons requires both Gαq/11 and Gβγ. J Neurosci. 2000;20:5623–5629. doi: 10.1523/JNEUROSCI.20-15-05623.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 191.Hu C, Depuy SD, Yao J, McIntire WE, Barrett PQ. Protein kinase A activity controls the regulation of T-type CaV3.2 channels by Gβγ dimers. J Biol Chem. 2009;284:7465–7473. doi: 10.1074/jbc.M808049200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 192.Wolfe JT, Wang H, Howard J, Garrison JC, Barrett PQ. T-type calcium channel regulation by specific G-protein βγ subunits. Nature. 2003;424:209–213. doi: 10.1038/nature01772. [DOI] [PubMed] [Google Scholar]

Articles from Channels are provided here courtesy of Taylor & Francis

RESOURCES