Abstract
Neural M-type (KCNQ/Kv7) K+ channels control somatic excitability, bursting and neurotransmitter release throughout the nervous system. Their activity is regulated by multiple signalling pathways. In superior cervical ganglion sympathetic neurons, muscarinic M1, angiotensin II AT1, bradykinin B2 and purinergic P2Y agonists suppress M current (im). Probes of PLC activity show agonists of all four receptors to induce robust PIP2 hydrolysis. We have grouped these receptors into two related modes of action. One mode involves depletion of phosphatidylinositol 4,5-bisphosphate (PIP2) in the membrane, whose interaction with the channels is thought necessary for their function. The other involves IP3-mediated intracellular Ca2+ signals that stimulate PIP2 synthesis, preventing its depletion, and suppress im via calmodulin. Carbon-fibre amperometry can evaluate the effect of M channel activity on release of neurotransmitter. Consistent with the dominant role of M current in control of neuronal discharge, M channel openers, or blockers, reduced or augmented the evoked release of noradrenaline neurotransmitter from superior cervical ganglion (SCG) neurons, respectively. We seek to localize the subdomains on the channels critical to their regulation by PIP2. Based on single-channel recordings from chimeras between high-PIP2 affinity KCNQ3 and low-PIP2 affinity KCNQ4 channels, we focus on a 57-residue domain within the carboxy-terminus that is a possible PIP2 binding site. Homology modelling of this domain using the published structure of IRK1 channels as a template predicts a structure very similar to an analogous region in IRK1 channels, and shows a cluster of basic residues in the KCNQ2 domain to correspond to those implicated in PIP2 regulation of Kir channels. We discuss some important issues dealing with these topics.
Over two decades have passed since Brown and colleagues described a neuronal K+ current in bullfrog sympathetic cells that they named ‘M current’ due to its suppression by stimulation of muscarinic acetylcholine receptors (mAChRs) (Brown & Adams, 1980; Constanti & Brown, 1981). This time- and voltage-dependent K+ current has a threshold for activation at typical neuronal resting potentials, with greater activity upon depolarization. This characteristic and its lack of inactivation give M current a major impact on neuronal excitability. Suppression of M current by the activation of appropriate receptors or by pharmacological blockade allows neurons to fire more rapidly due to the reduction in accommodation or spike frequency adaptation (Jones et al. 1995; Wang et al. 1998; Peretz et al. 2005; Gamper et al. 2006; Yue & Yaari, 2006; Zaika et al. 2006, 2007). Underlain by the KCNQ family of genes (Wang et al. 1998), M channels, also known as Kv7 (Gutman et al. 2003), are localized throughout the nervous system (Wang et al. 1998; Cooper et al. 2001; Roche et al. 2002; Shah et al. 2002; Devaux et al. 2004; Pan et al. 2006), where the channels are inhibited by stimulation of a variety of Gq/11-coupled neurotransmitter receptors. The transduction mechanism between the receptors and channels has been the subject of intense investigation, and work over the past five years has shed considerable light on this topic (Delmas & Brown, 2005). This report focuses on our recent findings in the modes of inhibition of M-type channels in sympathetic neurons of the superior cervical ganglion (SCG), as well as the physiological roles these signalling pathways play.
Receptor specificity
Much effort has gone into understanding how specificity is achieved in cellular signalling. One paradigm invokes compartmentalization of signalling proteins within membrane microdomains (Kreienkamp, 2002; Delmas et al. 2004). Another hypothesizes local production or degradation of intracellular messengers that results in their local concentration gradients (reviewed in Gamper & Shapiro, 2007b). We consider PIP2 hydrolysis by PLC as the first step in signalling by Gq/11-coupled receptors. This process can now be monitored in real-time in individual living cells using probes such as the green fluorescent protein (GFP)-tagged PH domain of PLCδ (PLCδ-PH) (Stauffer et al. 1998; Raucher et al. 2000). This probe binds to both PIP2 and IP3. In unstimulated cells (Fig. 1C, left), cytoplasmic [IP3] is very low and there is much membrane PIP2, to which the probe localizes. Upon stimulation of PLC-linked receptors, the probe translocates to the IP3 accumulating in the cytoplasm (Fig. 1C, right), which can be optically monitored. In SCG neurons, the responses of PIP2 probes to muscarinic and bradykinin stimulation have been found to be equally strong (Gamper et al. 2004; Winks et al. 2005; Hughes et al. 2007). To compare them to those from other Gq/11-coupled receptors that act on M channels, such as the purinergic P2Y and angiotensin AT1 types, we performed confocal microscopy on cells stimulated with multiple agonists. Figure 1A shows such an experiment on a SCG neuron transfected with PLCδ-PH using the biolistic ‘gene gun’ method (Gamper & Shapiro, 2006). Sequential stimulation of P2Y, bradykinin B2 and muscarinic M1 receptors by their respective agonists induces a reversible translocation of the probe from membrane to cytoplasm, indicating strong PIP2 hydrolysis by all three. It is important to note that such translocations cannot be easily used to report actual depletion of membrane PIP2 since the probe binds to IP3 with an affinity > 10-fold greater than that of PIP2 (Hirose et al. 1999); instead, careful IP3 titration experiments are required (Xu et al. 2003; Winks et al. 2005). Figure 1B summarizes the increases in cytoplasmic fluorescence from the probe induced by each agonist. Note that the translocation produced by bradykinin is the strongest, although other experiments discussed below indicate that stimulation of B2 receptors does not appreciably deplete PIP2 in these cells. What these results do indicate is that stimulation of M1, AT1, B2 and P2Y receptors results in signalling mechanisms that have in common the activation of PLC and vigorous PIP2 hydrolysis.
Figure 1. The PLCδ-PH probe reports PIP2 hydrolysis upon stimulation of multiple Gq/11-coupled receptors.
A, confocal images from a SCG neuron sequentially stimulated by UTP, bradykinin (BK), and oxotremorine (Oxo-M). In the unstimulated cell (control), the probe was concentrated mainly in the membrane, bound to PIP2. Upon stimulation by agonist, activation of PLC hydrolyses PIP2 and the probe translocates to the cytoplasm, bound to IP3. B, summarized data of the average fluorescence intensity, F, in a cytoplasmic region, normalized to the average intensity over 30 s before agonist application, Fo, quantified as the percentage increase in F (ΔF/Fo) induced by the four agonists. C, a schematic depiction of movement of the PLCδ-PH probe before (left panel) and after (right panel) activation of PLC via stimulation Gq/11-coupled receptors. D, an example of M current regulation under whole-cell clamp before (a, c) and after (b, d) application of either bradykinin or Oxo-M, and after complete blockade after application of linopirdine (e). GPCR, G protein coupled receptor; PLC, phospholipase C; DAG, diacylglycerol; BK or brady, bradykinin; Oxo-M or oxo, oxotremorine-M; LP, linopirdine. (G. P. Tolstykh, O. Zaika and M.S. Shapiro, unpublished observations.)
Figure 1D shows an example of M current suppression in an SCG cell studied under whole-cell clamp by the muscarinic agonist oxotremorine and by bradykinin. Although PIP2 hydrolysis by stimulation of both M1 and B2 receptors must produce much cytoplasmic IP3, the former does not induce IP3-mediated rises in intracellular Ca2+ ([Ca2+]i) (Wanke et al. 1987; Beech et al. 1991; Shapiro et al. 1994; Delmas et al. 2002; Zaika et al. 2007). In contrast, stimulation of bradykinin B2 receptors does induce [Ca2+]i transients, and bradykinin suppression of M current requires IP3-mediated rises in cytoplasmic Ca2+. Thus, clamping of intracellular Ca2+, blockade of IP3 receptors and depletion of internal Ca2+ stores all severely attenuate M current suppression by bradykinin, but all have little effect on the muscarinic action (Cruzblanca et al. 1998; Bofill-Cardona et al. 2000). In SCG cells, purinergic P2Y and B2 receptors have seemed to act similarly (Bofill-Cardona et al. 2000) and we have recently shown over-expression of an IP3 scavenger or an IP3 5-phosphatase, both of which prevent IP3 accumulation, to selectively block the purinergic or bradykinin inhibitions of M current, but to spare the muscarinic one (Zaika et al. 2007). We have further shown calmodulin (CaM) to be involved in the signal involving IP3-mediated [Ca2+]i rises, acting as the Ca2+ sensor of M-type channels. Thus, expression in SCG neurons of a dominant-negative CaM that cannot bind Ca2+ (Geiser et al. 1991) makes native M current mostly insensitive to bradykinin or purinergic stimulation, but again spares the muscarinic action (Gamper & Shapiro, 2003; Gamper et al. 2005; Zaika et al. 2007).
The receptor-specific [Ca2+]i signals have important implications for phosphoinositide (PI) signalling, since not only can hormonal stimulation hydrolyse PIP2, it can also strongly stimulate PIP2 synthesis by acceleration of PI kinases (Loew, 2007). One such pathway involves the Ca2+-binding protein neuronal Ca2+ sensor-1 (NCS-1), a signalling molecule with a wide spectrum of actions (Burgoyne & Weiss, 2001), including up-regulation of PI4-kinase, the first step in PIP2 synthesis from PI (Zhao et al. 2001; Taverna et al. 2002; Koizumi et al. 2002; Rajebhosale et al. 2003). Here, our thinking on M current signalling is fertilized by our work on modulation of N-type Ca2+ channels of SCG neurons by Gq/11-coupled receptors. As for P/Q-type channels (Wu et al. 2002), N-type channels also require membrane PIP2 to be functional (Gamper et al. 2004). Whereas the suppression of the Ca2+ current by muscarinic or angiotensin II stimulation is robust (Hille, 1994), it is little suppressed by stimulation of bradykinin or purinergic receptors (Gamper et al. 2004; Zaika et al. 2007). We hypothesized that this receptor-specific modulation of Ca2+ channels is due to receptor-specific depletion of PIP2. In this scenario, concurrent stimulation of PIP2 synthesis by bradykinin and purinergic agonists via Ca2+/NCS-1-mediated acceleration of PI4-kinase activity compensates for consumption of PIP2 by PLC. Indeed, over-expression of a DN NCS-1 that cannot bind Ca2+ bestowed onto bradykinin and purinergic agonists the effect of Ca2+ current suppression (Gamper et al. 2004; Zaika et al. 2007), consistent with PIP2 levels now falling in the absence of compensatory PIP2 synthesis (Delmas et al. 2005). Evidence for NCS-1 action on PIP2 production comes from other over-expression studies in SCG neurons (Winks et al. 2005), neuroendocrine PC12 cells (Koizumi et al. 2002; Rajebhosale et al. 2003), chromaffin cells (Pan et al. 2002) and in brain synaptosomes (Zheng et al. 2005). All of these studies are consistent with hormonal stimulation of PIP2 synthesis via signals, and subsequent amplification of PI turnover.
Although this hypothesis is attractive, it raises several crucial issues. As for much in the lipid kinase field, the control of a PI4-kinase by an orthologue of mammalian NCS-1 was discovered in yeast, where the respective proteins are Pik1 and Frq1 (frequenin in Drosophila) (Hendricks et al. 1999). The mammalian orthologue of yeast Pik1 is PI4-kinase IIIβ, one of four isoforms of mammalian PI4-kinases cloned so far, and this is the isoform shown to be stimulated by NCS-1 (Balla & Balla, 2006). Several studies have localized the bulk of wortmannin-sensitive PI4-kinases IIIα and IIIβ to ER and golgi membranes (Wong et al. 1997; Taverna et al. 2002; Balla & Balla, 2006), not to the plasma membrane where replenishment of PIP2 pools would seem to have most relevance for channel modulation. Moreover, PI4-kinase IIIα has been suggested to be necessary for the replenishment of plasma membrane PIP2, not the PI4-kinase IIIβ whose activity is known to be augmented by NCS-1 (Balla et al. 2005, 2008). Clearly, the possible regulation of PI4-kinase IIIα as well by NCS-1 needs to be carefully explored. Furthermore, the large literature documenting the control of plasma membrane PIP2 abundance by wortmannin indicates either greater localization there of type III PI4-kinases than the current literature suggest, or intriguingly, the rapid transfer of PI(4)P from ER to plasma membranes, perhaps at spots of apposition of the two membranes such at those thought intrinsic to ICrac made by the Stim/Orai complex (Liou et al. 2005; Wu et al. 2006). Likewise, the clear effects of NCS-1 on plasma membrane levels of PIP and PIP2 are hard to understand without invoking regulation of PI4-kinases. A mechanism may be the rapid translocation of NCS-1/type III PI4-kinases to the membrane upon activation of PLC and rises in [Ca2+]i (Taverna et al. 2002). Or rather, the kinase may translocate to NCS-1 prelocalized to the membrane via its myristoylation, a feature necessary for its association with, and up-regulation of, PI4-kinases (Zhao et al. 2001; Rajebhosale et al. 2003; Zheng et al. 2005).
Our scheme for the dual modulatory pathways acting on M channels is summarized in Fig. 2. It envisions two basic types of Gq/11-mediated signalling mechanisms in SCG neurons. Both involve phospholipase C (PLC) and are underpinned by the need of M channels for membrane phosphatidylinositol 4,5-bisphospate (PIP2) to function (Suh & Hille, 2002; Ford et al. 2003; Zhang et al. 2003; Li et al. 2005; Suh et al. 2006). The first mechanism is used by M1 mACh and angiotensin II AT1 receptors and principally involves depletion of PIP2 (Zaika et al. 2006; Suh & Hille, 2007). Perhaps due to the lack of spatial colocalization with IP3 receptors (Delmas & Brown, 2002), stimulation of these receptors does not elicit [Ca2+]i rises, membrane PIP2 abundance is allowed to fall and the M channels are inhibited. The second mechanism is based on the inositol trisphosphate (IP3)-mediated rises in intracellular Ca2+ induced by bradykinin (BK) B2 and purinergic P2Y receptors (Cruzblanca et al. 1998; Bofill-Cardona et al. 2000; Delmas et al. 2002; Zaika et al. 2007) and subsequent Ca2+ binding to calmodulin (CaM), which acts on the channels (Gamper & Shapiro, 2003; Gamper et al. 2005; Zaika et al. 2007). The signals produced by these receptors also augment PI4-kinase activity via NCS-1, thereby stabilizing PIP2 levels. The Ca2+/CaM action could involve reduction in the channels' affinity for PIP2, which would then unbind from the channel proteins. Such competitive or allosteric regulation of the affinity of membrane transport proteins for regulatory PIP2 molecules as a mechanism for modulation is widespread, and may act as a coincidence-detector motif for spatiotemporal targeting of receptor stimulation to the proper ion channel targets within the cell (Gamper & Shapiro, 2007a,b; Logothetis et al. 2007b; Lopes et al. 2007).
Figure 2. Modes of M channel regulation in sympathetic neurons.
Shown are mechanisms of inhibition of M channels used by two types of Gq/11-coupled receptors in the SCG. Both types activate PLCβ, which in turn hydrolyses PIP2 to IP3 and diacylglycerol. The first is used by M1 muscarinic acetylcholine and AT1 angiotensin II receptors (left). These agonists are ineffective in producing cytoplasmic Ca2+ signals, probably because the receptors are too far away from ER IP3 receptors, and the IP3 produced dissipates away (thick red arrow). Thus, much PIP2 is consumed, PIP2 unbinds from M channels down the [PIP2] gradient, and M channels are suppressed. The second is used by bradykinin B2 and purinergic P2Y6 receptors (right). Due to the spatial colocalization of these receptors to ER IP3 receptors, cytoplasmic Ca2+ signals are elicited. The released Ca2+ binds to neuronal Ca2+ sensor-1 (NCS-1) and to calmodulin (CaM). NCS-1 promotes PIP2 synthesis via acceleration of PI4-kinase, providing PIP2 to the membrane (purple arrows), and stabilizing PIP2 levels in the face of PLC activity. CaM binds to carboxy-terminal domains of the channel and likely acts by reducing its affinity for PIP2, which then unbinds from the channel since tonic [PIP2] is now insufficient to maintain association with the channel, and M current is suppressed. PLC, phospholipase C; PIP2, phosphatidylinositol 4,5-bisphosphate; CaM, calmodulin; PI4-kinase, phosphoinositide 4-kinase; NCS-1, neuronal calcium sensor-1 protein; IP3R, inositol trisphosphate receptor; ER, endoplasmic reticulum.
Neurotransmitter release: presynaptic modulation?
Given the established role for M channels in regulation of action potential firing, it would seem to follow that M current should also control the release of neurotransmitter at release sites. This question can be easily investigated in sympathetic neurons that release noradrenaline (NA) transmitter, which is oxidized at a suitable carbon fibre electrode (CFE) held at a potential greater than the chemical potential of NA (Zhou & Misler, 1995). Figure 3A is a schematic diagram showing the amperometry technique, in which a CFE held at +600 mV is placed proximal to a cluster of SCG neurons (Koh, 2006). A stimulating electrode is placed against the bundle of processes emanating from the cells and a shock train applied sufficient to elicit action potentials at the somas. The spikes in the amperometric record report the exocytosis of vesicles very near the electrode, and the baseline reflects the NA concentration in the solution near the electrode. The area under the amperometric records reflects the total amount of NA released by the shock train. As shown in Fig. 3B, application of the specific M channel opener retigabine (10 μm) blunted the evoked release of NA, whereas Fig. 3C shows that application of the specific M channel blocker XE991 (10 μm) augmented it. Such experiments are summarized in Fig. 3D, showing the significant and reversible effects on evoked release by the M channel specific compounds. These data are consistent with the regulation of neurotransmitter release by M channel activity, likely via the control of action-potential firing in response to excitatory inputs. We predict that neurotransmitters which modulate M channels will likewise regulate presynaptic release, although the issue is made complex by the concurrent modulation of Ca2+ channels by stimulation of several of these same Gq/11-coupled receptors. Indeed, stimulation of NA receptors in SCG neurons, which inhibits Ca2+ channels, but not M current, has been shown via this same method to strongly reduce the evoked release of NA (Koh & Hille, 1997).
Figure 3. M channels affect the release of NA from cultured sympathetic neurons detected by carbon-fibre amperometry.
A, schematic illustration of experimental conditions. A carbon-fibre electrode (CFE) is placed proximal to a cluster of SCG neurons, and the cells excited by 18 1 ms shocks from an extracellular field electrode (e-stim). The applied shocks induce many action potentials, causing the release of NA, both from the soma, and from varicosities in the neurites that are symbolized as small spots. A multibarrelled solution exchange system for perfusion of experimental solutions completes the set-up. In control conditions (B and C), the train of shocks elicited a total amperometric current that is proportional to the concentration of NA. The effects on NA release of retigabine (B) and XE991 (C) are shown. Consistent with the effects of these compounds on somatic excitability and action potential firing, retigabine caused a strong decrease in released NA, and XE991 caused an increase in released NA, suggesting that the effects of altering M channel activity on neuronal excitability may be paralleled by their effects on neurotransmitter release at nerve terminals. D, bars show the summarized data from these experiments. In the presence of retigabine, the train of shocks evoked a total charge of 68 ± 22% (P < 0.05) of control, and after XE991, the train of shocks evoked a total charge of 148 ± 19% (P < 0.05) of control. The effects of both drugs were fully reversible (O. Zaika and M. S. Shapiro, unpublished observations.)
Much accumulating work shows the role of somatodendritic M channels in control of excitability and neuronal discharge in a variety of peripheral and central neurons (Jones et al. 1997; Passmore et al. 2003; Yue & Yaari, 2004; Gu et al. 2005; Peters et al. 2005; Shen et al. 2005; Lawrence et al. 2006; Otto et al. 2006; Vervaeke et al. 2006; Wladyka & Kunze, 2006; Yue & Yaari, 2006; Zaika et al. 2006, 2007), which must indirectly regulate release of neurotransmitter. However, is the regulation of neurotransmitter release by M channels only through control of action-potential firing, or is there a direct role at the nerve terminal? The experimental evidence on this question is promising. Presynaptic M channels have been suggested to regulate neurotransmitter release of loaded [3H]noradrenaline, [3H]GABA and d-[3H]aspartate from hippocampal synaptosomes (Martire et al. 2004). In hippocampal slices, the M channel opener NH6 blunted spontaneous spiking behaviour and reduced the frequency of spontaneous EPSCs in cultured neurons, as well as evoked discharge, whereas the M channel blocker linopirdine had the opposite effect (Peretz et al. 2007). The simplest explanation for these presynaptic inhibitory actions would be M channel activation that hyperpolarizes the presynaptic terminals, thereby directly or indirectly depressing vesicular transmitter release. This notion is supported by simultaneous ionic current and Ca2+ measurements at the calyx of Held that show an exquisite dependence of neurotransmitter release on the resting potential in the nerve terminal, via tuning of tonic [Ca2+]i (Awatramani et al. 2005).
In sympathetic ganglia, preganglionic axons synapse onto the somatodendritic region of the primary ganglionic neurons that we routinely study. It is clear in those cells that suppression of M current alters cell excitability and spike-frequency adaptation (Brown, 1988; Jones et al. 1995; Zaika et al. 2006, 2007). Thus, it would seem that M1, AT1, B2 and P2Y receptors would all increase evoked NA release from sympathetic neurons in vitro (Kubista & Boehm, 2006). The results reported here using CFE amperometry and M channel specific drugs are consistent with the link between M current activity and NA release. However, CFE experiments that assayed the effect of stimulation of receptors that suppress the Ca2+ currents in SCG neurons indicated that those acting on Ca2+ channels alone, and not M channels, uniformly caused an inhibition of NA release, whereas those that concurrently suppress both types of channels, such as M1 receptors, had little effect. One explanation may be compensatory excitatory and inhibitory effects on neurotransmitter release by these two types of channels (Koh & Hille, 1997). Indeed, AT1 receptor stimulation, which also suppresses both types of channels (Shapiro et al. 1994; Zaika et al. 2006), also does not affect NA release (Gobel et al. 2000; our unpublished observations) whereas stimulation of B2 bradykinin and P2Y6 purinergic receptors, which inhibit M channels, but not Ca2+ channels, results in robust augmentation of NA release (Bofill-Cardona et al. 2000; Edelbauer et al. 2005). Since these effects on NA release are prevented by use of the M channel opener retigabine (Lechner et al. 2003), the actions are mediated via control of M channel activity. It will be interesting to rigorously assay the end-result on neurotransmitter release of receptor stimulation which acts on both types of channels, which may play different roles in regulating synaptic strength.
PIP2 sensitivity
With the role of PIP2 in the regulation of M channels established, our lab is currently seeking to determine the site(s) of action of PIP2 on the channel proteins, and the loci that determine PIP2 efficacy. Compared to other Kv channels, the C terminus of KCNQ channels is relatively extended, and has been shown to contain the sites of action for several regulatory molecules such as Ca2+/CaM, protein kinase C and A-kinase anchoring protein 79/150, and circumstantial evidence suggests PIP2 acts there as well (Delmas & Brown, 2005). Using recombinant KCNQ subunits studied in single-channel patches, we have found these channels to display divergent open probabilities arising from differential PIP2 sensitivities (Li et al. 2004, 2005). Based on the existence of highly basic clusters in the C-termini of the channels suggestive of PIP2-binding domains (Rosenhouse-Dantsker & Logothetis, 2007), as well as the PIP2 affinity shift caused by neutralization of a histidine residue located very proximal in the C-terminus of KCNQ2 (Zhang et al. 2003), we hypothesize that the C-terminus contains the molecular determinants of PIP2 regulation. To address this question, we are taking advantage of the highly differential PIP2 apparent affinities of KCNQ2–4 to construct KCNQ3/4 chimeras containing differing amounts of the C-terminal tail, and are mutating positively charged amino acids in a linker located between the first and second regions of highly conserved sequences in the C-terminus. The chimera data suggest this linker at least partly determines PIP2 apparent affinity, and is a possible binding site. Point mutants within this linker in KCNQ2 and KCNQ3 further localize the critical region to a cluster of basic residues (unpublished observations).
To gain further insight into possible interactions between PIP2 and KCNQ channels, a homology model of the C-terminal domain of KCNQ2 was constructed using the solved crystal structure of the cytoplasmic domain of IRK1 (Kir2.1) (Pegan et al. 2005) as a template (Protein Database accession number 1u4f), using SWISS-MODEL (Schwede et al. 2003). Since no crystal structure of the possible modulatory regions of KCNQ C-termini is available, we aligned the 57-residue core (aa 428–484) of our putative PIP2-binding domain of KCNQ2 to residues 186–245 in IRK, which has been identified as a PIP2 binding locus in those channels (Zhang et al. 1999). Figure 4A shows a homology model of the linker domain of KCNQ2 (blue) superimposed with the homologous region of IRK1 (red), indicating a marked structural similarity between the two with a CαRMS (carbon α root mean-square) of 0.55 Å. Although such a homology model is highly speculative, it may serve to guide more specific mutagenesis. Figure 4B depicts the two domains individually, with certain basic residues indicated. In the KCNQ2 linker, we are investigating the possible involvement of basic residues K452, R459, R461, R463 and R467 (Fig. 4B, right), which may play similar roles to R189, R218, K219 and R228 (Fig. 4B, left) previously identified in IRK1 as critical for PIP2 action (Logothetis et al. 2007a). We hypothesize this domain of KCNQ channels to be a locus of PIP2 binding.
Figure 4. Homology modelling of a putative PIP2-binding domain within the C terminus of KCNQ2 shows structural similarity to a similar domain in IRK1.
Shown are the results of homology modelling using the program SWISS-MODEL. A, superimposed are the structure of residues 186–245 of IRK1 (Kir2.1) and the predicted structure of residues 428–484 of KCNQ2 using the IRK1 structure as a template. B, shown individually are the homology model of the linker domain of KCNQ2 (blue) and the homologous region of IRK1 (red), with elements of secondary structure indicated. Shown are clusters of basic residues (stick model) located in a putative PIP2 binding site on IRK1 (Logothetis et al. 2007a), as well as a cluster of basic residues located in the C-terminal domain of KCNQ currently under investigation as critical for PIP2 regulation. (C. C. Hernandez and M. S. Shapiro, unpublished.)
Conclusions
Given the identification of the KCNQ2 and KCNQ3 gene products that underlie many M channels (Wang et al. 1998) from inherited human epileptic syndromes (Biervert et al. 1998; Charlier et al. 1998; Singh et al. 1998; Lerche et al. 2001), it is unsurprising that these channels are under widespread study for new therapeutic modes for a variety of diseases. Thus, specific M channel openers, such as retigabine, are currently being developed as novel antiepileptic drugs (Cooper & Jan, 2003) and specific blockers, such as XE991 (a linopirdine derivative) (Zaczek et al. 1998) may find use as ‘cognition enhancers’. Since M current plays such a powerful role in tuning synaptic efficacy and neurotransmitter action, we also predict M channels to be potential targets to ameliorate a variety of psychiatric diseases, as well as new treatments for chronic pains, in accord with the identification of M currents in a variety of nociceptive/sensory neurons (Passmore et al. 2003; Wladyka & Kunze, 2006). We look forward to the expanding literature that will indicate the diverse physiological roles of M channels, the molecules that regulate them, and their multifaceted modulatory pathways.
Acknowledgments
We thank Pamela Martin for expert technical assistance and Luis Gimenez for scientific discussions. This work was supported by NIH grant R01 NS043394 to M.S.S.
References
- Awatramani GB, Price GD, Trussell LO. Modulation of transmitter release by presynaptic resting potential and background calcium levels. Neuron. 2005;48:109–121. doi: 10.1016/j.neuron.2005.08.038. [DOI] [PubMed] [Google Scholar]
- Balla A, Balla T. Phosphatidylinositol 4-kinases: old enzymes with emerging functions. Trends Cell Biol. 2006;16:351–361. doi: 10.1016/j.tcb.2006.05.003. [DOI] [PubMed] [Google Scholar]
- Balla A, Ju Kim Y, Varnai P, Szentpetery Z, Knight Z, Shokat KM, Balla T. Maintenance of hormone-sensitive phosphoinositide pools in the plasma membrane requires phosphatidylinositol 4-kinase IIIα. Mol Biol Cell. 2008 doi: 10.1091/mbc.E07-07-0713. in press. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Balla A, Tuymetova G, Tsiomenko A, Varnai P, Balla T. A plasma membrane pool of phosphatidylinositol 4-phosphate is generated by phosphatidylinositol 4-kinase type-III alpha: studies with the PH domains of the oxysterol binding protein and FAPP1. Mol Biol Cell. 2005;16:1282–1295. doi: 10.1091/mbc.E04-07-0578. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Beech DJ, Bernheim L, Mathie A, Hille B. Intracellular Ca2+ buffers disrupt muscarinic suppression of Ca2+ current and M current in rat sympathetic neurons. Proc Natl Acad Sci U S A. 1991;88:652–656. doi: 10.1073/pnas.88.2.652. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Biervert C, Schroeder BC, Kubisch C, Berkovic SF, Propping P, Jentsch TJ, Steinlein OK. A potassium channel mutation in neonatal human epilepsy. Science. 1998;279:403–406. doi: 10.1126/science.279.5349.403. [DOI] [PubMed] [Google Scholar]
- Bofill-Cardona E, Vartian N, Nanoff C, Freissmuth M, Boehm S. Two different signaling mechanisms involved in the excitation of rat sympathetic neurons by uridine nucleotides. Mol Pharmacol. 2000;57:1165–1172. [PubMed] [Google Scholar]
- Brown DA. M Currents. In: Narahashi T, editor. Ion Channels. New York: Plenum; 1988. pp. 55–94. [DOI] [PubMed] [Google Scholar]
- Brown DA, Adams PR. Muscarinic suppression of a novel voltage-sensitive K+ current in a vertebrate neurone. Nature. 1980;283:673–676. doi: 10.1038/283673a0. [DOI] [PubMed] [Google Scholar]
- Burgoyne RD, Weiss JL. The neuronal calcium sensor family of Ca2+-binding proteins. Biochem J. 2001;353:1–12. [PMC free article] [PubMed] [Google Scholar]
- Charlier C, Singh NA, Ryan SG, Lewis TB, Reus BE, Leach RJ, Leppert M. A pore mutation in a novel KQT-like potassium channel gene in an idiopathic epilepsy family. Nat Genet. 1998;18:53–55. doi: 10.1038/ng0198-53. [DOI] [PubMed] [Google Scholar]
- Constanti A, Brown DA. M-Currents in voltage-clamped mammalian sympathetic neurones. Neurosci Lett. 1981;24:289–294. doi: 10.1016/0304-3940(81)90173-7. [DOI] [PubMed] [Google Scholar]
- Cooper EC, Harrington E, Jan YN, Jan LY. M channel KCNQ2 subunits are localized to key sites for control of neuronal network oscillations and synchronization in mouse brain. J Neurosci. 2001;21:9529–9540. doi: 10.1523/JNEUROSCI.21-24-09529.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cooper EC, Jan LY. M-channels: neurological diseases, neuromodulation, and drug development. Arch Neurol. 2003;60:496–500. doi: 10.1001/archneur.60.4.496. [DOI] [PubMed] [Google Scholar]
- Cruzblanca H, Koh DS, Hille B. Bradykinin inhibits M current via phospholipase C and Ca2+ release from IP3-sensitive Ca2+ stores in rat sympathetic neurons. Proc Natl Acad Sci U S A. 1998;95:7151–7156. doi: 10.1073/pnas.95.12.7151. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Delmas P, Brown DA. Junctional signaling microdomains: bridging the gap between the neuronal cell surface and Ca2+ stores. Neuron. 2002;36:787–790. doi: 10.1016/s0896-6273(02)01097-8. [DOI] [PubMed] [Google Scholar]
- Delmas P, Brown DA. Pathways modulating neural KCNQ/M (Kv7) potassium channels. Nat Rev Neurosci. 2005;6:850–862. doi: 10.1038/nrn1785. [DOI] [PubMed] [Google Scholar]
- Delmas P, Coste B, Gamper N, Shapiro MS. Phosphoinositide lipid second messengers: new paradigms for calcium channel modulation. Neuron. 2005;47:179–182. doi: 10.1016/j.neuron.2005.07.001. [DOI] [PubMed] [Google Scholar]
- Delmas P, Crest M, Brown DA. Functional organization of PLC signaling microdomains in neurons. Trends Neurosci. 2004;27:41–47. doi: 10.1016/j.tins.2003.10.013. [DOI] [PubMed] [Google Scholar]
- Delmas P, Wanaverbecq N, Abogadie FC, Mistry M, Brown DA. Signaling microdomains define the specificity of receptor-mediated InsP3 pathways in neurons. Neuron. 2002;34:209–220. doi: 10.1016/s0896-6273(02)00641-4. [DOI] [PubMed] [Google Scholar]
- Devaux JJ, Kleopa KA, Cooper EC, Scherer SS. KCNQ2 is a nodal K+ channel. J Neurosci. 2004;24:1236–1244. doi: 10.1523/JNEUROSCI.4512-03.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Edelbauer H, Lechner SG, Mayer M, Scholze T, Boehm S. Presynaptic inhibition of transmitter release from rat sympathetic neurons by bradykinin. J Neurochem. 2005;93:1110–1121. doi: 10.1111/j.1471-4159.2005.03084.x. [DOI] [PubMed] [Google Scholar]
- Ford CP, Stemkowski PL, Light PE, Smith PA. Experiments to test the role of phosphatidylinositol 4,5-bisphosphate in neurotransmitter-induced M-channel closure in bullfrog sympathetic neurons. J Neurosci. 2003;23:4931–4941. doi: 10.1523/JNEUROSCI.23-12-04931.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gamper N, Li Y, Shapiro MS. Structural requirements for differential sensitivity of KCNQ K+ channels to modulation by Ca2+/calmodulin. Mol Biol Cell. 2005;16:3538–3551. doi: 10.1091/mbc.E04-09-0849. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gamper N, Reznikov V, Yamada Y, Yang J, Shapiro MS. Phosphatidylinositol 4,5-bisphosphate signals underlie receptor-specific Gq/11-mediated modulation of N-type Ca2+ channels. J Neurosci. 2004;24:10980–10992. doi: 10.1523/JNEUROSCI.3869-04.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gamper N, Shapiro MS. Calmodulin mediates Ca2+-dependent modulation of M-type K+ channels. J Gen Physiol. 2003;122:17–31. doi: 10.1085/jgp.200208783. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gamper N, Shapiro MS. Exogenous expression of proteins in neurons using the biolistic particle delivery system. Methods Mol Biol. 2006;337:27–38. doi: 10.1385/1-59745-095-2:27. [DOI] [PubMed] [Google Scholar]
- Gamper N, Shapiro MS. Regulation of ion transport proteins by membrane phosphoinositides. Nat Rev Neurosci. 2007a;8:921–934. doi: 10.1038/nrn2257. [DOI] [PubMed] [Google Scholar]
- Gamper N, Shapiro MS. Target-specific PIP2 signalling: how might it work? J Physiol. 2007b;582:967–975. doi: 10.1113/jphysiol.2007.132787. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gamper N, Zaika O, Li Y, Martin P, Hernandez CC, Perez MR, Wang AY, Jaffe DB, Shapiro MS. Oxidative modification of M-type K+ channels as a mechanism of cytoprotective neuronal silencing. EMBO J. 2006;25:4996–5004. doi: 10.1038/sj.emboj.7601374. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Geiser JR, van Tuinen D, Brockerhoff SE, Neff MM, Davis TN. Can calmodulin function without binding calcium? Cell. 1991;65:949–959. doi: 10.1016/0092-8674(91)90547-c. [DOI] [PubMed] [Google Scholar]
- Gobel I, Trendelenburg AU, Cox SL, Meyer A, Starke K. Electrically evoked release of [3H]noradrenaline from mouse cultured sympathetic neurons: release-modulating heteroreceptors. J Neurochem. 2000;75:2087–2094. doi: 10.1046/j.1471-4159.2000.0752087.x. [DOI] [PubMed] [Google Scholar]
- Gu N, Vervaeke K, Hu H, Storm JF. Kv7/KCNQ/M and HCN/h, but not KCa2/SK channels, contribute to the somatic medium after-hyperpolarization and excitability control in CA1 hippocampal pyramidal cells. J Physiol. 2005;566:689–715. doi: 10.1113/jphysiol.2005.086835. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gutman GA, Chandy KG, Adelman JP, Aiyar J, Bayliss DA, Clapham DE, Covarriubias M, Desir GV, Furuichi K, Ganetzky B, Garcia ML, Grissmer S, Jan LY, Karschin A, Kim D, Kuperschmidt S, Kurachi Y, Lazdunski M, Lesage F, Lester HA, McKinnon D, Nichols CG, O'Kelly I, Robbins J, Robertson GA, Rudy B, Sanguinetti M, Seino S, Stuehmer W, Tamkun MM, Vandenberg CA, Wei A, Wulff H, Wymore RS. International Union of Pharmacology. XLI. Compendium of voltage-gated ion channels: Potassium channels. Pharmacol Rev. 2003;55:583–586. doi: 10.1124/pr.55.4.9. [DOI] [PubMed] [Google Scholar]
- Hendricks KB, Wang BQ, Schnieders EA, Thorner J. Yeast homologue of neuronal frequenin is a regulator of phosphatidylinositol-4-OH kinase. Nat Cell Biol. 1999;1:234–241. doi: 10.1038/12058. [DOI] [PubMed] [Google Scholar]
- Hille B. Modulation of ion-channel function by G-protein-coupled receptors. Trends Neurosci. 1994;17:531–536. doi: 10.1016/0166-2236(94)90157-0. [DOI] [PubMed] [Google Scholar]
- Hirose K, Kadowaki S, Tanabe M, Takeshima H, Iino M. Spatiotemporal dynamics of inositol 1,4,5-trisphosphate that underlies complex Ca2+ mobilization patterns. Science. 1999;284:1527–1530. doi: 10.1126/science.284.5419.1527. [DOI] [PubMed] [Google Scholar]
- Hughes S, Marsh S, Tinker A, Brown D. PIP2-dependent inhibition of M-type (Kv7.2/7.3) potassium channels: direct on-line assessment of PIP2 depletion by Gq-coupled receptors in single living neurons. Pflugers Arch. 2007;455:115–124. doi: 10.1007/s00424-007-0259-6. [DOI] [PubMed] [Google Scholar]
- Jones S, Brown DA, Milligan G, Willer E, Buckley NJ, Caulfield MP. Bradykinin excites rat sympathetic neurons by inhibition of M current through a mechanism involving B2 receptors and Gαq/11. Neuron. 1995;14:399–405. doi: 10.1016/0896-6273(95)90295-3. [DOI] [PubMed] [Google Scholar]
- Jones LP, Patil PG, Snutch TP, Yue DT. G-protein modulation of N-type calcium channel gating current in human embryonic kidney cells (HEK 293) J Physiol. 1997;498:601–610. doi: 10.1113/jphysiol.1997.sp021886. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Koh D-S. Carbon fiber amperometry in the study of ion channels and secretion. In: Stockard J, Shapiro M, editors. Ion Channels: Methods and Protocols. Totowa, NJ, USA: Humana Press; 2006. pp. 139–153. [DOI] [PubMed] [Google Scholar]
- Koh DS, Hille B. Modulation by neurotransmitters of catecholamine secretion from sympathetic ganglion neurons detected by amperometry. Proc Natl Acad Sci U S A. 1997;94:1506–1511. doi: 10.1073/pnas.94.4.1506. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Koizumi S, Rosa P, Willars GB, Challiss RA, Taverna E, Francolini M, Bootman MD, Lipp P, Inoue K, Roder J, Jeromin A. Mechanisms underlying the neuronal calcium sensor-1-evoked enhancement of exocytosis in PC12 cells. J Biol Chem. 2002;277:30315–30324. doi: 10.1074/jbc.M201132200. [DOI] [PubMed] [Google Scholar]
- Kreienkamp HJ. Organisation of G-protein-coupled receptor signalling complexes by scaffolding proteins. Curr Opin Pharmacol. 2002;2:581–586. doi: 10.1016/s1471-4892(02)00203-5. [DOI] [PubMed] [Google Scholar]
- Kubista H, Boehm S. Molecular mechanisms underlying the modulation of exocytotic noradrenaline release via presynaptic receptors. Pharmacol Ther. 2006;112:213–242. doi: 10.1016/j.pharmthera.2006.04.005. [DOI] [PubMed] [Google Scholar]
- Lawrence JJ, Saraga F, Churchill JF, Statland JM, Travis KE, Skinner FK, McBain CJ. Somatodendritic Kv7/KCNQ/M channels control interspike interval in hippocampal interneurons. J Neurosci. 2006;26:12325–12338. doi: 10.1523/JNEUROSCI.3521-06.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lechner SG, Mayer M, Boehm S. Activation of M1 muscarinic receptors triggers transmitter release from rat sympathetic neurons through an inhibition of M-type K+ channels. J Physiol. 2003;553:789–802. doi: 10.1113/jphysiol.2003.052449. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lerche H, Jurkat-Rott K, Lehmann-Horn F. Ion channels and epilepsy. Am J Med Genet. 2001;106:146–159. doi: 10.1002/ajmg.1582. [DOI] [PubMed] [Google Scholar]
- Li Y, Gamper N, Hilgemann DW, Shapiro MS. Regulation of Kv7 (KCNQ) K+ channel open probability by phosphatidylinositol 4,5-bisphosphate. J Neurosci. 2005;25:9825–9835. doi: 10.1523/JNEUROSCI.2597-05.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li Y, Gamper N, Shapiro MS. Single-channel analysis of KCNQ K+ channels reveals the mechanism of augmentation by a cysteine-modifying reagent. J Neurosci. 2004;24:5079–5090. doi: 10.1523/JNEUROSCI.0882-04.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Liou J, Kim ML, Heo WD, Jones JT, Myers JW, Ferrell JE, Jr, Meyer T. STIM is a Ca2+ sensor essential for Ca2+-store-depletion-triggered Ca2+ influx. Curr Biol. 2005;15:1235–1241. doi: 10.1016/j.cub.2005.05.055. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Loew LM. Where does all the PIP2 come from? J Physiol. 2007;582:945–951. doi: 10.1113/jphysiol.2007.132860. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Logothetis DE, Jin T, Lupyan D, Rosenhouse-Dantsker A. Phosphoinositide-mediated gating of inwardly rectifying K+ channels. Pflugers Arch. 2007a;455:83–95. doi: 10.1007/s00424-007-0276-5. [DOI] [PubMed] [Google Scholar]
- Logothetis DE, Lupyan D, Rosenhouse-Dantsker A. Diverse Kir modulators act in close proximity to residues implicated in phosphoinositide binding. J Physiol. 2007b;582:953–965. doi: 10.1113/jphysiol.2007.133157. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lopes CM, Remon JI, Matavel A, Sui JL, Keselman I, Medei E, Shen Y, Rosenhouse-Dantsker A, Rohacs T, Logothetis DE. Protein kinase A modulates PLC-dependent regulation and PIP2-sensitivity of K+ channels. Channels. 2007;1:124–134. doi: 10.4161/chan.4322. [DOI] [PubMed] [Google Scholar]
- Martire M, Castaldo P, D'Amico M, Preziosi P, Annunziato L, Taglialatela M. M channels containing KCNQ2 subunits modulate norepinephrine, aspartate, and GABA release from hippocampal nerve terminals. J Neurosci. 2004;24:592–597. doi: 10.1523/JNEUROSCI.3143-03.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Otto JF, Yang Y, Frankel WN, White HS, Wilcox KS. A spontaneous mutation involving Kcnq2 (Kv7.2) reduces M-current density and spike frequency adaptation in mouse CA1 neurons. J Neurosci. 2006;26:2053–2059. doi: 10.1523/JNEUROSCI.1575-05.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pan CY, Jeromin A, Lundstrom K, Yoo SH, Roder J, Fox AP. Alterations in exocytosis induced by neuronal Ca2+ sensor-1 in bovine chromaffin cells. J Neurosci. 2002;22:2427–2433. doi: 10.1523/JNEUROSCI.22-07-02427.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pan Z, Kao T, Horvath Z, Lemos J, Sul JY, Cranstoun SD, Bennett V, Scherer SS, Cooper EC. A common ankyrin-G-based mechanism retains KCNQ and NaV channels at electrically active domains of the axon. J Neurosci. 2006;26:2599–2613. doi: 10.1523/JNEUROSCI.4314-05.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Passmore GM, Selyanko AA, Mistry M, Al-Qatari M, Marsh SJ, Matthews EA, Dickenson AH, Brown TA, Burbidge SA, Main M, Brown DA. KCNQ/M currents in sensory neurons: significance for pain therapy. J Neurosci. 2003;23:7227–7236. doi: 10.1523/JNEUROSCI.23-18-07227.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pegan S, Arrabit C, Zhou W, Kwiatkowski W, Collins A, Slesinger PA, Choe S. Cytoplasmic domain structures of Kir2.1 and Kir3.1 show sites for modulating gating and rectification. Nat Neurosci. 2005;8:279–287. doi: 10.1038/nn1411. [DOI] [PubMed] [Google Scholar]
- Peretz A, Degani N, Nachman R, Uziyel Y, Gibor G, Shabat D, Attali B. Meclofenamic acid and diclofenac, novel templates of KCNQ2/Q3 potassium channel openers, depress cortical neuron activity and exhibit anticonvulsant properties. Mol Pharmacol. 2005;67:1053–1066. doi: 10.1124/mol.104.007112. [DOI] [PubMed] [Google Scholar]
- Peretz A, Sheinin A, Yue C, Degani-Katzav N, Gibor G, Nachman R, Gopin A, Tam E, Shabat D, Yaari Y, Attali B. Pre- and postsynaptic activation of M-channels by a novel opener dampens neuronal firing and transmitter release. J Neurophysiol. 2007;97:283–295. doi: 10.1152/jn.00634.2006. [DOI] [PubMed] [Google Scholar]
- Peters HC, Hu H, Pongs O, Storm JF, Isbrandt D. Conditional transgenic suppression of M channels in mouse brain reveals functions in neuronal excitability, resonance and behavior. Nat Neurosci. 2005;8:51–60. doi: 10.1038/nn1375. [DOI] [PubMed] [Google Scholar]
- Rajebhosale M, Greenwood S, Vidugiriene J, Jeromin A, Hilfiker S. Phosphatidylinositol 4-OH kinase is a downstream target of neuronal calcium sensor-1 in enhancing exocytosis in neuroendocrine cells. J Biol Chem. 2003;278:6075–6084. doi: 10.1074/jbc.M204702200. [DOI] [PubMed] [Google Scholar]
- Raucher D, Stauffer T, Chen W, Shen K, Guo S, York JD, Sheetz MP, Meyer T. Phosphatidylinositol 4,5-bisphosphate functions as a second messenger that regulates cytoskeleton-plasma membrane adhesion. Cell. 2000;100:221–228. doi: 10.1016/s0092-8674(00)81560-3. [DOI] [PubMed] [Google Scholar]
- Roche JP, Westenbroek R, Sorom AJ, Hille B, Mackie K, Shapiro MS. Antibodies and a cysteine-modifying reagent show correspondence of M current in neurons to KCNQ2 and KCNQ3 K+ channels. Br J Pharmacol. 2002;137:1173–1186. doi: 10.1038/sj.bjp.0704989. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rosenhouse-Dantsker A, Logothetis DE. Molecular characteristics of phosphoinositide binding. Pflugers Arch. 2007;455:45–53. doi: 10.1007/s00424-007-0291-6. [DOI] [PubMed] [Google Scholar]
- Schwede T, Kopp J, Guex N, Peitsch MC. SWISS-MODEL: An automated protein homology-modeling server. Nucleic Acids Res. 2003;31:3381–3385. doi: 10.1093/nar/gkg520. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Shah MM, Mistry M, Marsh SJ, Brown DA, Delmas P. Molecular correlates of the M-current in cultured rat hippocampal neurons. J Physiol. 2002;544:29–37. doi: 10.1113/jphysiol.2002.028571. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Shapiro MS, Wollmuth LP, Hille B. Angiotensin II inhibits calcium and M current channels in rat sympathetic neurons via G proteins. Neuron. 1994;12:1319–1329. doi: 10.1016/0896-6273(94)90447-2. [DOI] [PubMed] [Google Scholar]
- Shen W, Hamilton SE, Nathanson NM, Surmeier DJ. Cholinergic suppression of KCNQ channel currents enhances excitability of striatal medium spiny neurons. J Neurosci. 2005;25:7449–7458. doi: 10.1523/JNEUROSCI.1381-05.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Singh NA, Charlier C, Stauffer D, DuPont BR, Leach RJ, Melis R, Ronen GM, Bjerre I, Quattlebaum T, Murphy JV, McHarg ML, Gagnon D, Rosales TO, Peiffer A, Anderson VE, Leppert M. A novel potassium channel gene, KCNQ2, is mutated in an inherited epilepsy of newborns. Nat Genet. 1998;18:25–29. doi: 10.1038/ng0198-25. [DOI] [PubMed] [Google Scholar]
- Stauffer TP, Ahn S, Meyer T. Receptor-induced transient reduction in plasma membrane PtdIns(4,5)P2 concentration monitored in living cells. Curr Biol. 1998;8:343–346. doi: 10.1016/s0960-9822(98)70135-6. [DOI] [PubMed] [Google Scholar]
- Suh B, Hille B. Recovery from muscarinic modulation of M current channels requires phosphatidylinositol 4,5-bisphosphate synthesis. Neuron. 2002;35:507–520. doi: 10.1016/s0896-6273(02)00790-0. [DOI] [PubMed] [Google Scholar]
- Suh BC, Hille B. Regulation of KCNQ channels by manipulation of phosphoinositides. J Physiol. 2007;582:911–916. doi: 10.1113/jphysiol.2007.132647. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Suh BC, Inoue T, Meyer T, Hille B. Rapid chemically induced changes of PtdIns(4,5)P2 gate KCNQ ion channels. Science. 2006;314:1454–1457. doi: 10.1126/science.1131163. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Taverna E, Francolini M, Jeromin A, Hilfiker S, Roder J, Rosa P. Neuronal calcium sensor 1 and phosphatidylinositol 4-OH kinase β interact in neuronal cells and are translocated to membranes during nucleotide-evoked exocytosis. J Cell Sci. 2002;115:3909–3922. doi: 10.1242/jcs.00072. [DOI] [PubMed] [Google Scholar]
- Vervaeke K, Gu N, Agdestein C, Hu H, Storm JF. Kv7/KCNQ/M-channels in rat glutamatergic hippocampal axons and their role in regulation of excitability and transmitter release. J Physiol. 2006;576:235–256. doi: 10.1113/jphysiol.2006.111336. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang HS, Pan Z, Shi W, Brown BS, Wymore RS, Cohen IS, Dixon JE, McKinnon D. KCNQ2 and KCNQ3 potassium channel subunits: molecular correlates of the M-channel. Science. 1998;282:1890–1893. doi: 10.1126/science.282.5395.1890. [DOI] [PubMed] [Google Scholar]
- Wanke E, Ferroni A, Malgaroli A, Ambrosini A, Pozzan T, Meldolesi J. Activation of a muscarinic receptor selectively inhibits a rapidly inactivated Ca2+ current in rat sympathetic neurons. Proc Natl Acad Sci U S A. 1987;84:4313–4317. doi: 10.1073/pnas.84.12.4313. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Winks JS, Hughes S, Filippov AK, Tatulian L, Abogadie FC, Brown DA, Marsh SJ. Relationship between membrane phosphatidylinositol-4,5-bisphosphate and receptor-mediated inhibition of native neuronal M channels. J Neurosci. 2005;25:3400–3413. doi: 10.1523/JNEUROSCI.3231-04.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wladyka CL, Kunze DL. KCNQ/M-currents contribute to the resting membrane potential in rat visceral sensory neurons. J Physiol. 2006;575:175–189. doi: 10.1113/jphysiol.2006.113308. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wong K, Meyers ddR, Cantley LC. Subcellular locations of phosphatidylinositol 4-kinase isoforms. J Biol Chem. 1997;272:13236–13241. doi: 10.1074/jbc.272.20.13236. [DOI] [PubMed] [Google Scholar]
- Wu L, Bauer CS, Zhen XG, Xie C, Yang J. Dual regulation of voltage-gated calcium channels by PtdIns(4,5)P2. Nature. 2002;419:947–952. doi: 10.1038/nature01118. [DOI] [PubMed] [Google Scholar]
- Wu MM, Buchanan J, Luik RM, Lewis RS. Ca2+ store depletion causes STIM1 to accumulate in ER regions closely associated with the plasma membrane. J Cell Biol. 2006;174:803–813. doi: 10.1083/jcb.200604014. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Xu C, Watras J, Loew LM. Kinetic analysis of receptor-activated phosphoinositide turnover. J Cell Biol. 2003;161:779–791. doi: 10.1083/jcb.200301070. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yue C, Yaari Y. KCNQ/M channels control spike afterdepolarization and burst generation in hippocampal neurons. J Neurosci. 2004;24:4614–4624. doi: 10.1523/JNEUROSCI.0765-04.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yue C, Yaari Y. Axo-somatic and apical dendritic Kv7/M channels differentially regulate the intrinsic excitability of adult rat CA1 pyramidal cells. J Neurophysiol. 2006;95:3480–3495. doi: 10.1152/jn.01333.2005. [DOI] [PubMed] [Google Scholar]
- Zaczek R, Chorvat RJ, Saye JA, Pierdomenico ME, Maciag CM, Logue AR, Fisher BN, Rominger DH, Earl RA. Two new potent neurotransmitter release enhancers, 10,10-bis(4-pyridinylmethyl)-9(10H)-anthracenone and 10,10-bis(2-fluoro-4-pyridinylmethyl)-9(10H)-anthracenone: comparison to linopirdine. J Pharmacol Exp Ther. 1998;285:724–730. [PubMed] [Google Scholar]
- Zaika O, Lara LS, Gamper N, Hilgemann DW, Jaffe DB, Shapiro MS. Angiotensin II regulates neuronal excitability via phosphatidylinositol 4,5-bisphosphate-dependent modulation of Kv7 (M-type) K+ channels. J Physiol. 2006;575:49–67. doi: 10.1113/jphysiol.2006.114074. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zaika O, Tolstykh GP, Jaffe DB, Shapiro MS. Inositol triphosphate-mediated Ca2+ signals direct purinergic P2Y-receptor regulation of neuronal ion channels. J Neurosci. 2007;27:8914–8926. doi: 10.1523/JNEUROSCI.1739-07.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhang H, Craciun LC, Mirshahi T, Rohacs T, Lopes CM, Jin T, Logothetis DE. PIP2 activates KCNQ channels, and its hydrolysis underlies receptor-mediated inhibition of M currents. Neuron. 2003;37:963–975. doi: 10.1016/s0896-6273(03)00125-9. [DOI] [PubMed] [Google Scholar]
- Zhang H, He C, Yan X, Mirshahi T, Logothetis DE. Activation of inwardly rectifying K+ channels by distinct PtdIns(4,5)P2 interactions. Nat Cell Biol. 1999;1:183–188. doi: 10.1038/11103. [DOI] [PubMed] [Google Scholar]
- Zhao X, Varnai P, Tuymetova G, Balla A, Toth ZE, Oker-Blom C, Roder J, Jeromin A, Balla T. Interaction of neuronal calcium sensor-1 (NCS-1) with phosphatidylinositol 4-kinase β stimulates lipid kinase activity and affects membrane trafficking in COS-7 cells. J Biol Chem. 2001;276:40183–40189. doi: 10.1074/jbc.M104048200. [DOI] [PubMed] [Google Scholar]
- Zheng Q, Bobich JA, Vidugiriene J, McFadden SC, Thomas F, Roder J, Jeromin A. Neuronal calcium sensor-1 facilitates neuronal exocytosis through phosphatidylinositol 4-kinase. J Neurochem. 2005;92:442–451. doi: 10.1111/j.1471-4159.2004.02897.x. [DOI] [PubMed] [Google Scholar]
- Zhou Z, Misler S. Amperometric detection of stimulus-induced quantal release of catecholamines from cultured superior cervical ganglion neurons. Proc Natl Acad Sci U S A. 1995;92:6938–6942. doi: 10.1073/pnas.92.15.6938. [DOI] [PMC free article] [PubMed] [Google Scholar]




