Skip to main content
The Journal of Physiology logoLink to The Journal of Physiology
. 2008 May 8;586(Pt 13):3043–3054. doi: 10.1113/jphysiol.2008.153460

Ca2+ microdomains near plasma membrane Ca2+ channels: impact on cell function

Anant B Parekh 1
PMCID: PMC2538792  PMID: 18467365

Abstract

In eukaryotic cells, a rise in cytoplasmic Ca2+ can activate a plethora of responses that operate on time scales ranging from milliseconds to days. Inherent to the use of a promiscuous signal like Ca2+ is the problem of specificity: how can Ca2+ activate some responses but not others? We now know that the spatial profile of the Ca2+ signal is important Ca2+ does not simply rise uniformly throughout the cytoplasm upon stimulation but can reach very high levels locally, creating spatial gradients. The most fundamental local Ca2+ signal is the Ca2+ microdomain that develops rapidly near open plasmalemmal Ca2+ channels like voltage-gated L-type (Cav1.2) and store-operated CRAC channels. Recent work has revealed that Ca2+ microdomains arising from these channels are remarkably versatile in triggering a range of responses that differ enormously in both temporal and spatial profile. Here, I delineate basic features of Ca2+ microdomains and then describe how these highly local signals are used by Ca2+-permeable channels to drive cellular responses.

Introduction

Eukaryotic cells respond to changes in their environment by generating intracellular signals. Foremost amongst such signals is the Ca2+ ion. Although its appreciation by physiologists was rather fortuitous, beginning with Sidney Ringer's serendipitous discovery that it was Ca2+ in London's tap water that supported frog heart contractions, we now know that a rise in cytoplasmic Ca2+ can activate a plethora of key cellular responses including neurotransmitter release, metabolism, cell growth and proliferation as well as cell death (Berridge et al. 2003; Clapham, 2008). This remarkable versatility of Ca2+ raises the old chestnut of specificity: how can such a promiscuous messenger ever elicit a selective response? For many years, this problem could be described by Churchill's aphorism of ‘a riddle wrapped in a mystery inside an enigma’. Over the past few years, however, evidence has revealed that information is contained in both the frequency and the amplitude of Ca2+ signals, and these discrete components can be deciphered by intracellular proteins and then translated into distinct responses (Clapham, 2008). There is now a growing awareness that local Ca2+ signals or Ca2+ microdomains, generated by open Ca2+ channels, provide a further means for generating specificity (Neher, 1998b; Rizzuto & Pozzan, 2006). Diffusion of Ca2+ through an open Ca2+ channel creates a local Ca2+ gradient, which falls off steeply with distance due to cytoplasmic Ca2+ buffering. Within a few nanometres of the channel pore, buffers are simply not fast enough to capture the incoming Ca2+ (Neher, 1998a,b). This, coupled with the very small volume the microdomain occupies, means that local Ca2+ can rapidly rise to very high levels, often orders of magnitude greater than the bulk cytoplasmic Ca2+. Location of targets close to Ca2+ channels therefore provides a means for rapid and selective activation (Berridge et al. 2003). The term Ca2+ microdomain is widely used in the literature to refer to a range of Ca2+ signals from local plumes of Ca2+ near open single Ca2+ channels to those restricted to one pole or region of a cell and occupying distances of hundreds of nanometres to micrometres. In this review, I will focus exclusively on how Ca2+ microdomains arising from plasma membrane Ca2+ channels affect cell function, drawing mainly from two very well-studied systems: voltage-gated L-type Ca2+ channels in neurons and store-operated CRAC channels in non-excitable cells. Ca2+ microdomains near Ca2+ release channels on intracellular stores play important roles in generating inter alia intracellular Ca2+ oscillations, Ca2+ wave propagation and mitochondrial ATP production, but are not discussed here. The reader is referred instead to two excellent recent reviews on this topic (Petersen, 2005; Rizzuto & Pozzan, 2006).

Measuring and Manipulating Ca2+ microdomains

The amplitude and spatial extent of local Ca2+ influx ([Ca2+]i through a Ca2+ channel is determined by the single channel Ca2+ flux and, at distances greater than a few nanometres, local Ca2+ buffering (Neher, 1998a,b). Ca2+ flux in turn is determined by the electrochemical gradient for Ca2+ entry and the single channel conductance.

In the absence of buffers, the spatial profile of the Ca2+ signal is determined by simple diffusion and can be described thus (Neher, 1998a):

graphic file with name tjp0586-3043-m1.jpg

where, iCa is the single channel Ca2+ current, DCa is the diffusion coefficient of Ca2+ in the cytoplasm (300 μm2 s−1), r the distance from the channel and [Ca2+]o the bulk Ca2+ concentration.

In living cells, mobile and immobile buffers shape the profile of the Ca2+ signal. Immobile buffers are fixed in space and hence will only prolong the time needed to reach steady-state, and not affect the final level reached. Once bound with Ca2+, immobile buffers can be ignored as they are not replaced by free buffer. Mobile buffers on the other hand bind Ca2+ and then, by buffered diffusion, spread the Ca2+ deeper into the cytoplasm. Ca2+-bound mobile buffers are replaced by Ca2+-free ones and thus will bind Ca2+ until the total pool is saturated.

Equation (1) can then be modified to take into account the presence of excess unsaturable mobile buffers (Neher, 1998a):

graphic file with name tjp0586-3043-m2.jpg

where λ = √DCa/konB and denotes the mean path length a Ca2+ ion travels before being captured by chelator, kon is the chelator's on-rate and B is the free buffer concentration where buffer is present in excess.

Unfortunately, we still know little about the kon, concentration and spatial distribution of mobile Ca2+ buffers in cells. Moreover, different cell types have different endogenous Ca2+ binding capacities (Neher, 1995), and hence it is difficult to compare directly the spatial profile of a Ca2+ signal through different Ca2+ channels in different systems. In addition, most single channel recordings of voltage-gated Ca2+ channels have understandably been conducted with 110 mm Ba2+ as the charge carrier, not physiological Ca2+. Two channels where the conductance has been studied in physiological Ca2+ are L-type Ca2+ channels (2.6 pS) and store-operated CRAC channels (unitary chord conductance of 9 fS) (Zweifach & Lewis, 1993; Church & Stanley, 1996). The profile of the Ca2+ microdomain as a function of distance from each of these channels is simulated in Fig. 1A and B. Ca2+ close to the pore is considerably higher than bulk Ca2+ and even a Ca2+-selective but low conductance channel like CRAC can nevertheless generate quite a high local Ca2+ concentration.

Figure 1. Amplitude and spatial extent of Ca2+ microdomains.

Figure 1

A and B, plots of the numerically calculated profile of a Ca2+ microdomain near an open L-type Ca2+ channel (A) and a CRAC channel (B). Single channel currents were 0.144 pA (Church & Stanley, 1996) and 1.5 fA (Zweifach & Lewis, 1995), respectively. Endogenous Ca2+ buffering was ignored. C, the distance (λ) a Ca2+ ion diffuses before it is captured by some widely used Ca2+ chelators is shown and is stated next to each chelator. All concentrations are in mm. D, various properties of widely used fluorescent dyes (Lattanzio. & Bartschat (1991) and the calculated λ. All concentrations were 100 μm, corresponding to the levels used experimentally. For C and D, resting Ca2+ was set at 100 nm and the free chelator concentration then calculated. λ is not a precise calculation because buffer is not present in excess.

The simulations in Fig. 1 neglect Ca2+ buffering. How might this sculpt the profile of the Ca2+ signal? The mean path length (λ) Ca2+ diffuses from a point source for different concentrations of experimentally used chelators is shown in Fig. 1C. Strong spatial restriction of the microdomain is seen with millimolar BAPTA whereas EGTA is much less effective, reflecting its ∼100-fold slower kon. Figure 1C also shows that the fluorescent dye fura 2, widely used to measure cytoplasmic Ca2+, has a marked effect on the lateral spread of a Ca2+ microdomain. This reflects its rapid kon (7.6 × 108m−1 s−1; Lattanzio. & Bartschat, 1991), which is similar to that of BAPTA. Figure 1D presents a similar analysis for commonly used fluorescent Ca2+ indicators. All reduce the lateral spread of single microdomains, which needs to be borne in mind when interpreting Ca2+ signals measured with such dyes, especially when used in high concentrations. These calculations are somewhat simplistic as they do not correct for buffered diffusion. Nor do they consider local saturation of buffer, although this would be a less acute problem for CRAC channels where local Ca2+ is severalfold less than free buffer concentration.

Numerical calculations therefore predict high local Ca2+ concentrations near single channels. Can such local Ca2+ be measured? Early work comparing the rate and extent of activation of bio-assays such as Ca2+-activated K+ channels (Roberts et al. 1990; Robitaille et al. 1990; Prakriya et al. 1996) or exocytosis (Heidelberger et al. 1994; Beutner et al. 2004) following a bulk cytoplasmic Ca2+ rise with activation of voltage-gated Ca2+ channels enabled researchers to conclude that local Ca2+ near the Ca2+ channels could reach 10–100 μm. Can these microdomains be measured directly? Unfortunately, this has proved very difficult because the resolution of light microscopy is not sufficiently high to detect the Ca2+ microdomain near a single open plasma membrane Ca2+ channel (see section on Neurotransmitter release). An exciting new development is the use of total internal reflection fluorescence microscopy to measure flux through single channels (Demuro & Parker, 2004), although this method does not permit combined patch clamp and optical recordings of the channel under investigation.

There are several ways to invoke the involvement of local Ca2+ entry rather than a bulk Ca2+ rise in activating a target. Firstly, one can test whether manipulating the amplitude and spatial extent of single Ca2+ channel flux (by changing the driving force) alters the response without changing bulk Ca2+. Alternatively, for a similar bulk Ca2+, many open Ca2+ channels but each with a smaller single channel flux might be less effective in driving a response than fewer open channels but each with a larger single channel flux. Secondly, one can compare the effects of exogenous Ca2+ chelators, loaded into the cell, on the response. Very close to the channel pore, chelators have virtually no effect on the Ca2+ concentration but do reduce the lateral spread of the microdomain in a manner determined by the chelator's on-rate (Neher, 1998b). Use of chelators with different on-rates can therefore provide information on the effects of lateral spread of the microdomain on a response as well as providing an estimate of how close to the pore an effector is located. However, chelators have effects in addition to their ability to buffer Ca2+. For example, BAPTA can interfere with InsP3 binding to InsP3 receptors (Richardson & Taylor, 1993) and impedes protein kinase C activation (Chang et al. 2008). Hence effects of BAPTA are not always attributable to Ca2+ buffering and these additional effects need to be considered. Third, the kinetics of a response provides clues as to whether it is driven by adjacent Ca2+ channels. Fast responses arising within milliseconds, which therefore develop well before bulk Ca2+ has reached equilibrium, often point towards a role for local Ca2+ signals. Fourth, if the Ca2+ requirement of a process is considerably higher than the increase in bulk Ca2+ that is reached, then it is difficult to see how it can be controlled by anything other than local Ca2+. Finally, in some cases, it has been possible to show the Ca2+ channel and its target physically interact (see next section). Such strong colocalization implicates control by local Ca2+.

Ca2+ microdomains from voltage-gated Ca2+ channels and neuronal function

Activation of Ca2+-dependent K+ channels

Ca2+ microdomains near voltage-gated Ca2+ channels are tightly coupled to Ca2+-activated K+ channels in vestibular hair cells (Roberts et al. 1990), adrenal chromaffin cells (Prakriya et al. 1996) and the frog neuromuscular junction (Robitaille et al. 1990). A striking example of this tight coupling has been reported recently by Falker and colleagues (Berkefeld et al. 2006). Affinity purification of large conductance Ca2+-activated K+ (BKCa) channels from plasma membrane of rat brain revealed the additional presence of L-, P/Q- and N-type Ca2+ channels, suggesting these form a macromolecular complex. Giant inside-out patches from Xenopus oocytes expressing recombinant BK and P/Q-type channels revealed rapid activation of the K+ current by Ca2+ influx, consistent with a tight physical interaction (Berkefeld et al. 2006). Comparing BKCa kinetics following voltage-dependent Ca2+ influx with that seen upon directly varying cytoplasmic Ca2+ led the authors to conclude that local Ca2+ was ∼10 μm. Finally, high concentrations of cytoplasmic EGTA were ineffective, but BAPTA reduced BKCa–P/Q-type channel coupling. The distance between BKCa and P/Q-type channels was placed at ∼10–15 nm (Berkefeld et al. 2006). Although the physical basis for coupling is unclear, a major role for the pore forming α subunit was found.

The coupling between BK and P/Q-type channels is a nice example of a tight physical interaction between a Ca2+ channel and its target, thus ensuring robust activation of the effector. Other examples include the interaction between N-type channels at active zones with syntaxin 1, a protein that binds to synaptic vesicles and thus locates the vesicle near the channel, and coupling between L-type Ca2+ channels and ryanodine type 1 release channels of the sarcoplasmic reticulum at the t-tubules of skeletal muscle. In this latter case, coupling is conformational in nature and not by Ca2+ flux through the L-type channel.

Neurotransmitter release

At synapses, neurotransmitter release following an action potential clearly has to be fast in order to maximize information propagation. Hence one might expect a dominant role for local Ca2+ entry in driving the fusion of ready releasable vesicles. Imaging of n-aequorin-J, a luminescent protein with low affinity for Ca2+, revealed ‘hot-spots’ of subplasmalemmal Ca2+ following stimulation of presynaptic terminals of the squid giant synapse and which were ∼200–300 μm in amplitude (Llinas et al. 1992). Action potentials induced localized presynaptic subplasmalemmal Ca2+ hot-spots in Xenopus neuromuscular junction preparations and these fluorescent spots decayed within a couple of milliseconds (DiGregorio & Vergara, 1997). The spatial resolution with optical microscopy is limited by the wavelength of light. Even with a high numerical aperture quartz objective lens, it would be more than ∼200 nm, which is severalfold wider than the calculated spatial extent of a Ca2+ microdomain (Neher, 1998b). Using total internal reflection fluorescence microscopy, which measures events located within ∼100 nm of the plasma membrane, local ‘hot spots’ of high Ca2+ have been seen in several different experimental systems and these are likely to correlate with the fusion event (Becherer et al. 2003; Zenisek et al. 2003). Another approach has been to compare the rate and extent of exocytosis between opening of Ca2+ channels and a bulk rise in cytoplasmic Ca2+ (Heidelberger et al. 1994; Beutner et al. 2004). The best way to achieve a rapid and uniform rise in bulk Ca2+ is to use flash photolysis of caged Ca2+, since the Ca2+ sensor triggering exocytosis and the fluorescent dye used to measure this Ca2+ will be exposed to the same cytoplasmic Ca2+ load (Neher, 1998b). Using this approach, it has become clear that the Ca2+ concentration at the release sites is in the range 10–300 μm, depending on the type of neuron investigated (Heidelberger et al. 1994; Schneggenburger & Neher, 2000; Beutner et al. 2004; Bollmann & Sakmann, 2005). With these higher values in particular, one might expect secretion to be unaffected by slow chelators like EGTA. Even at tens of millimolar concentrations, EGTA had no effect on exoctyosis at the squid giant synapse whereas BAPTA inhibited release at considerably lower concentrations (Adler et al. 1991). Hence at this synapse, Ca2+ influx is very tightly coupled with transmitter release. On the other hand, high local Ca2+ does not necessarily mean that ready releasable vesicles are governed by single Ca2+ channel microdomains. Work in the calyx of Held and neocortex have shown that EGTA can reduce evoked neurotransmitter release (Borst & Sakmann, 1996; Rozov et al. 2001), demonstrating that Ca2+ channels are not as tightly coupled to exocytosis in these synapses as they are in the squid giant synapse. Instead, several Ca2+ channels are likely to affect the release of one vesicle at the calyx of Held. This is supported by a morphological and modelling study that reported an average distance of 200 nm between a vesicle and a Ca2+ channel and that a cluster of Ca2+ channels controlled the fusion of a vesicle (Meinrenken et al. 2002). Local Ca2+ is still the trigger, but a stringent colocalization between channel and effector is not required, at least in this synapse.

Neuronal growth

One key aspect of neuronal differentiation involves the targeting of axons to form synapses with other neurons, myocytes or endocrine cells. Such pathfinding occurs through migration of growth cones at the axon's tip (Spitzer, 2006). Growth cones exhibit spontaneous fluctuations in intracellular Ca2+ (around 1–10 spikes per hour). These Ca2+ signals are spatially restricted to the growth cones but are nevertheless needed for neuronal differentiation. Because these Ca2+ signals can occur in the absence of action potentials but depend on Ca2+ influx, the Ca2+ entry pathway is thought to involve an ion channel active at the resting potential. TRPC channels are Ca2+ permeable non-selective cation channels expressed in a range of diverse cell types (Nilius et al. 2007). Recent evidence implicates a major role for TRPC channels in growth cone morphology, motility and neurite elongation. In young hippocampal neurons, TRPC5 forms a complex with stathmin2, a protein that accumulates in growth cones, and this complex is transported to the cones via a vesicular transport mechanism (Greka et al. 2003). Whereas most growth cones had filipodia with perimeters (an index of length) of < 5 μm, overexpression of TRPC5 increased this ∼3-fold.

Growth cones do not simply extend along one axis but can turn in response to diffusible chemoattractants like netrin 1 and brain-derived neurotrophic factor (BDNF). Axon turning is an intracellular Ca2+-dependent process and the site of the Ca2+ rise is critical (Zheng, 2000). Focal laser-induced photolysis of caged Ca2+ has revealed that a local Ca2+ rise on one side of a growth cone causes the turning of the growth cone to that side (Zheng, 2000). Physiologically, these local Ca2+ rises can be brought about by opening of Ca2+ channels. In cultured Xenopus neurons, netrin 1 and BDNF both activated TRPC currents and knockdown of TRPC1 expression prevented both growth cone turning and the Ca2+ rise in response to nestrin-1 (Wang & Poo, 2005). In cultured cerebellar granule neurons, growth cone turning to BDNF was suppressed by knockdown of TRPC3 channels (Li et al. 2005). Collectively, these findings all point to a central role for TRPC channels in regulating growth cone motility and turning.

Is a microdomain of Ca2+ through a single TRPC channel sufficient to drive growth cone motility and turning? Zheng (2000) measured the Ca2+ rise to be ∼78 nm at the site of uncaging Ca2+ in the growth cone. Although TRPC channels are non-selective, conducting Ca2+ and Na+ with a permeability ratio of ∼1.6–9.0, they have a high single channel conductance of the order of 16–75 pS (Nilius et al. 2007). Hence the Ca2+ microdomain near an open TRPC channel may be sufficient to trigger the growth turning process.

Gene expression

In neurons, activation of the transcriptional factor CREB (cAMP response element binding protein) is linked to synaptic plasticity, learning and memory (Bailey et al. 1996; Shaywitz & Greenberg, 1999; Ahn et al. 2000). CREB activation is a two step process (Hu et al. 1999; Shaywitz & Greenberg, 1999): first, phosphorylation of serine 133 by either the ERK1/2 pathway or Ca2+/calmodulin-dependent kinases occurs and this is followed by binding of the CREB binding protein, a transcriptional coactivator. CREB binding protein in turn is activated by nuclear Ca2+ and CaM kinase IV.

A rise in nuceloplasmic Ca2+ following depolarization of the plasma membrane is sufficient to activate CREB (Hardingham et al. 1997). Direct injection of dextran-conjugated BAPTA into the nucleus reduced the nucleoplasmic Ca2+ rise and prevented activity-induced CREB stimulation without compromising the cytoplasmic Ca2+ elevation. Since cytoplasmic Ca2+ readily propagates into the nucleus, Ca2+ channel opening at the plasma membrane can raise nuclear Ca2+ by eliciting a bulk Ca2+ rise. In addition to this global Ca2+ dependence, a substantial body of evidence has demonstrated that local Ca2+ influx can also have an impact upon nuclear events. Ca2+ microdomains arising from voltage-gated Ca2+ channels (Deisseroth et al. 1996; 1998; Dolmetsch et al. 2001) or Ca2+-permeable NMDA receptors (Hardingham et al. 2001) can signal to the nucleus and induce CREB phosphorylation.

Although cortical and hippocampal neurons coexpress L-type, P/Q-type and N-type channels, only the L-type channel consistently couples to CREB activation (Deisseroth et al. 1998; Dolmetsch et al. 2001), providing a poignant example of how local Ca2+ signalling can achieve specificity.

The L-type Ca2+ channel has a calmodulin-binding isoleucine–glutamine (IQ) motif in the carboxy terminus (Zuehlke et al. 1999). At resting Ca2+ (∼100 nm), calmodulin is stably associated with this motif. Hence calmodulin is held in a strategically optimal location adjacent to the Ca2+ channel pore, enabling it to detect Ca2+ microdomains arising from L-type channel opening. Calmodulin translocates from the cytoplasm to the nucleus in a Ca2+-dependent manner, where it stimulates CREB phosphorylation through activation of nuclear Ca2+/calmodulin-dependent protein kinases (Deisseroth et al. 1998). In cortical neurons, the IQ motif is required for signalling to the ERK pathway and the subsequent activation of CREB (Dolmetsch et al. 2001). However, voltage-dependent Ca2+ channels like the P/Q-type channel also bind calmodulin (Lee et al. 1999) yet are less able to activate CREB (Dolmetsch et al. 2001). Hence additional domains of the L-type channel are likely to be involved in signalling to the nucleus. One such region is the PDZ interaction sequence located within the amino acid sequence valine–serine–asparagine–leucine (VSNL) at the end of the L-type channel (Weick et al. 2003). Overexpression of a GFP-tagged VSNL construct prevented L-type channels in hippocampal CA3–CA1 pyramidal neurons from evoking CREB phosphorlyation and subsequent gene transcription. Hence the L-type channel may be physically integrated into a signalling complex that efficiently links the local Ca2+ microdomain to a pathway culminating in CREB activation.

Recently, it has been found that the carboxy terminus of the L-type Ca2+ channel can be cleaved to release a peptide that translocates to the nucleus where it regulates expression of numerous genes including the gap junction Cx31.1, netrin 4, TRPV4 and the sodium–calcium exchanger (Gomez-Ospina et al. 2006). How the L-type channel is cleaved and whether this involves a protease activated by local Ca2+ remain unclear at present.

Can L-type channels distinguish between different types of excitatory stimulus? Synaptic activity (excitatory postsynaptic potentials) is much more effective than spike activity (back-propagating action potentials) in triggering CREB phosphorylation, despite both stimuli generating similar bulk Ca2+ elevations. This is probably because L-type channels behave as kinetic filters, activating slowly upon depolarization (Mermelstein et al. 2000). Moreover, they open at relatively hyperpolarized potentials. Both the voltage- and time-dependent gating of the channels permit greater Ca2+ flux during a slow excitatory postsynaptic potential than with a brief action potential, providing a mechanism whereby certain synaptic inputs can trigger more CREB phosphorylation than others.

L-type channels are not the only voltage-gated Ca2+ channels that signal to the nucleus though. In cerebellar granule cells, activation of P/Q-type channels resulted in increased transcription and translation of syntaxin-1A. L- and N-type channels were not able to mimic this (Sutton et al. 1999). Ca2+ influx through these P/Q-type channels was propagated to the nucleus via Ca2+-induced Ca2+ release. Presumably, there is a specific association between P/Q-type channels and the internal Ca2+ stores, which the other Ca2+ channels are excluded from. Whether this tight coupling involves local Ca2+ influx through the P/Q-type channels is not known.

Activation of NMDA receptors induces CREB activation and gene expression. As this is still manifest in the presence of voltage-gated Ca2+ channel blockers, it is Ca2+ influx through the NMDA receptors themselves drives gene expression. In EGTA-loaded hippocampal neurons, NMDA receptor stimulation still activated the ERK pathway leading to CREB phosphorylation. Hence Ca2+ microdomains around open NMDA channels are thought to provide the trigger for signal transmission to the nucleus (Hardingham et al. 2001). These local Ca2+ signals are apparently propagated to the nucleus as a regenerative Ca2+ wave involving intracellular stores, probably through a Ca2+-induced Ca2+ release mechanism.

Several NMDA receptor subtypes have been described in neurons and are composed of different subunits (Cull-Candy et al. 2001). Subunit composition imparts differences in Mg2+ block as well as fractional Ca2+ permeability. Moreover, Ca2+ flux through NMDA channels can be increased within seconds following phosphorylation by protein kinase A (Skeberdis et al. 2006) or the Src family of tyrosine kinases (Takasu et al. 2002). Since the amplitude and spatial extent of a microdomain depends on Ca2+ flux through the channel, it will be interesting to see whether the local Ca2+ signals emanating from different NMDA channels, or the same channel in the presence of active protein kinase A for example, differ in their ability to trigger CREB phosphorylation. This could provide a novel mechanism for regulating synaptic plasticity.

Not only is the Ca2+ flux through NMDA receptors important in regulating nuclear events, but the location of the channels is also critical. Ca2+ microdomains associated with synaptic NMDA receptors triggered CREB phosphorylation and BDNF expression but extra-synaptic NMDA receptors had the opposite effect, switching off the CREB pathway and evoking cell death (Hardingham et al. 2002). These differences were related to the different subunit composition of the channels.

Ca2+ microdomains near store-operated Ca2+ channels and non-excitable cell function

Store-operated Ca2+ channels are activated by the emptying of intracellular Ca2+ stores and the ensuing Ca2+ entry activates a disparate array of responses (Parekh & Putney, 2005; Parekh, 2007). The most thoroughly studied store-operated channel is the CRAC channel (Hoth & Penner, 1992; Parekh & Putney, 2005). The molecular basis of this ubiquitous pathway has recently been identified (reviewed in Lewis, 2007). STIM1 is the Ca2+ sensor that detects the fall in store Ca2+ content and then migrates to form discrete puncta just below the plasma membrane. Here it somehow opens Orai1, the CRAC channel pore. The reader is referred to other symposium articles in this issue for more discussion of these proteins.

Local Ca2+ entry and ion channels

In a prostate cancer cell line, store-operated Ca2+ influx partially inhibited a volume-regulated anion channel that was involved in regulatory volume decrease (Lemonnier et al. 2002). Dialysis with 1 μm free Ca2+ failed to mimic this effect, indicating that subplasmalemmal [Ca2+] after CRAC channel opening was higher than this. It was suggested that store-operated Ca2+ channels were colocalized with the Cl channels.

Another example was revealed in a detailed study on TRPC7 channels, which are activated by both store-operated and non-store-operated mechanisms, the latter involving diacylglycerol. Diacylglycerol activation of TRPC7 was suppressed by block of SERCA pumps (Lemonnier et al. 2006). It was found that Ca2+ entry through TRPC7 fed back to inhibit the channels but this was prevented by Ca2+ uptake into the stores. What was particularly striking was that interaction between TRPC7 and SERCA pumps was maintained in the presence of high concentrations of BAPTA, revealing an intimate interaction between these proteins that reside on different membranes (Lemonnier et al. 2006).

Ca2+ tunnelling: local refilling of the endoplasmic reticulum

A striking example of how local Ca2+ entry can impact upon a response some distance away has been dissected out from studies in pancreatic acinar cells. In these highly polarized epithelia, store-operated Ca2+ entry occurs across the basolateral membrane but exocytosis of zymogen-containing granules takes place at the apical pole several micrometres away. Simple diffusion of Ca2+ is impeded by the very high Ca2+ cytosolic buffering of the acinar cell (Mogami et al. 1999). This problem is circumvented by the ingenious mechanism of Ca2+ tunnelling (Mogami et al. 1997). The endoplasmic reticulum is a contiguous organelle in acinar cells and runs very close to the basolateral membrane as well as protruding into the apical pole (Petersen et al. 2001). Local Ca2+ entering across the basolateral membrane is taken up into the endoplasmic reticulum by SERCA pumps which are adjacent to the Ca2+ influx sites. The endoplasmic reticulum has a very low Ca2+ buffering capacity, so the Ca2+ can diffuse rapidly through the organelle to reach the apical region, where it is released close to the granules by InsP3 (Mogami et al. 1997). Local Ca2+ influx can therefore have an impact upon events restricted to the opposite pole of the cell.

Local Ca2+entry and regulation of adenylyl cyclases

The first real evidence that local Ca2+ entry through store-operated Ca2+ channels could regulate a cellular response was provided by Cooper and colleagues (reviewed in Willoughby & Cooper, 2007). They studied the dependence of calmodulin-regulated adenylyl cyclases on cytoplasmic Ca2+. Two isoforms of adenylyl cyclase are activated by a rise in cytoplasmic Ca2+ (types 1 and 8) and two are inhibited by a Ca2+ rise (types 5 and 6). These enzymes associate with the plasma membrane. In C6-2B glioma cells, robust Ca2+ release from intracellular stores in response to agonist, thapsigargin or the Ca2+ ionophore ionomycin were all unable to regulate the Ca2+-dependent isoforms (Chiono et al. 1995; Fagan et al. 1996). By contrast, store-operated Ca2+ influx significantly altered enzyme activity even though it raised bulk Ca2+ to a lesser extent than that achieved by Ca2+ release. Hence these isoforms are regulated by a subplasmalemmal Ca2+ rise rather than global Ca2+. However, these early results did not distinguish between a general subplasmalemmal Ca2+ rise or a more restricted Ca2+ signal in the vicinity of open store-operated Ca2+ channels. Further work by the Cooper laboratory clearly established the concept that local Ca2+ tightly associated with store-operated Ca2+ channels regulated adenylyl cyclase. Firstly, loading cells with BAPTA prevented store-operated entry from regulating adenylyl cyclase whereas EGTA was ineffective (Fagan et al. 1998). Second, the potency series of divalent cations for regulating adenylyl cyclase correlated with the conductivity through CRAC channels. When Ba2+ permeated store-operated channels, no regulation of adenylyl cyclase was detected whereas Sr2+ was ∼50% as effective as Ca2+ (Gu & Cooper, 2000). This is similar to the conductivity profile of CRAC channels, which over the physiological voltage range (–80 to 0 mV) is Ca2+ > Sr2+ > Ba2+, where Sr2+ conductivity is 50–80% that of Ca2+ (Zweifach & Lewis, 1993; Fierro & Parekh, 2000). By contrast, the divalent cation equilibrium binding profile to the EF hand of calmodulin is 1 Ca2+: 40 Sr2+: 1600 Ba2+. From equilibrium considerations, Sr2+ should therefore be much less effective than Ca2+ in regulating calmodulin-dependent adenylyl cyclases. The fact that Sr2+ is almost as effective as Ca2+ implies that adenylyl cyclase is exposed to very high local Sr2+ levels, which could only be the case if the enzyme was located very close to the Ca2+ channel. Finally, opening non-store-operated Ca2+ channels by application of either OAG (Martin & Cooper, 2006) or arachidonic acid (Shuttleworth & Thompson, 1999) did not affect adenylyl cyclase activity, despite raising bulk Ca2+ to a similar extent to that seen following store-operated entry. Hence only a local Ca2+ rise in the vicinity of store-operated channels can regulate adenylyl cyclase. What underpins this tight molecular coupling? One possibility is that Ca2+-dependent adenylyl cyclases are tethered to the store-operated channels, analogous to anchoring of protein kinase A to channels through AKAP proteins. Alternatively, store-operated channels and Ca2+-dependent adenylyl cyclases might associate with specific plasma membrane domains like lipid rafts (Willoughby & Cooper, 2007). Destruction of cholesterol-rich rafts by methyl-β-cyclodextrin prevented store-operated entry from regulating adenylyl cyclase without affecting the extent of Ca2+ influx, suggesting recruitment of both proteins to rafts facilitated interaction between them. Although certain TRPC channels have been identified in rafts, it is not clear whether CRAC channels are integrated here too. Now that Orai1 has been identified as the CRAC channel pore, this possibility can be tested directly.

Local Ca2+ entry and NO synthase

As with Ca2+-regulated adenylyl cyclases, eNOS is preferentially activated by local Ca2+ influx through store-operated channels (Lin et al. 2000). In endothelial cells, opening of store-operated channels following stimulation with thapsigargin in the presence of 4 mm external Ca2+ raised bulk cytoplasmic [Ca2+] to a level comparable to that seen with ionomycin in 30 μm external Ca2+. However, nitric oxide production was 10-fold higher with thapsigargin. Lin et al. (2000) then constructed two chimaeras: eNOS linked to aequorin and a myristoylation-deficient eNOS-aequorin. Myristoylation is required for eNOS insertion into the plasma membrane. The eNOS-aequorin construct associated with the plasma membrane whereas the myristoylation-deficient chimaera was cytosolic. Consistent with local Ca2+ influx and hence spatial Ca2+ gradients, eNOS-aequorin reported a significantly higher Ca2+ signal than the cytosolic probe following activation of store-operated Ca2+ entry.

Local Ca2+ entry and plasma membrane Ca2+ATPase pump

Another example of local Ca2+ influx tightly controlling a plasma membrane-restricted enzyme has come from work on T lymphocytes (Bautista & Lewis, 2004). A rise in cytoplasmic Ca2+ following CRAC channel opening increased plasmalemmal Ca2+-ATPase activity by reducing the KM and increasing Vmax of the pump. The increased Ca2+-ATPase activity was triggered by local Ca2+ microdomains near CRAC channels in addition to bulk cytoplasmic Ca2+ (Bautista & Lewis, 2004). Both an increase in the electrical gradient for Ca2+ influx by hyperpolarization of the membrane potential and raising the chemical gradient by elevating external Ca2+ significantly increased pump activity without changing bulk cytoplasmic Ca2+. Furthermore, a nice method was developed to reinforce the importance of microdomains in controlling effector activity. The trivalent cation La3+ is a potent blocker of CRAC channels, with an IC50 of ∼20 nm. La3+ is likely to inhibit CRAC channels in a way similar to that seen in voltage-gated Ca2+ channels, namely by slow permeation resulting in a flickery block (Lansman et al. 1986). La3+ does not reduce the driving force for Ca2+ entry and hence does not decrease the amplitude of the microdomain, though its duration is likely to be reduced. Bautista & Lewis (2004) found that bulk cytoplasmic Ca2+ following CRAC channel activation in 20 mm external Ca2+ and 7 nm La3+ was similar to that seen in 2 mm external Ca2+ without La3+. Despite this, pump activity was significantly enhanced in 20 mm Ca2+ and La3+, providing compelling evidence that Ca2+ microdomains near CRAC channels are the critical determinants of enhanced pump activity.

Local Ca2+ entry and signalling to cytoplasmic enzymes: insight from the cPLA2 and 5-lipoxygenase pathway

All the preceding examples of effectors activated by local Ca2+ influx are directly regulated by Ca2+ and are attached to, or span, the plasma membrane close to CRAC channels, thus facilitating their activation by local Ca2+ gradients. An interesting complication arises when the Ca2+-dependent target is located in the cytoplasm, some distance away. Recent work in my group has examined how local Ca2+ entry through CRAC channels in mast cells stimulates Ca2+-dependent phospholipase A2 (cPLA2) and 5-lipoxygenase, two cytoplasmic enzymes that generate the intracellular messenger arachidonic acid and the intercellular pro-inflammatory signal cysteinyl leukotriene LTC4, which is secreted from the cell. Aberrant LTC4 production is important clinically as it is linked to a variety of allergic diseases including allergic rhinitis and asthma (Peters-Golden et al. 2006).

Stimulation with thapsigargin or agonist in Ca2+-free solution elicited significant rises in global Ca2+ but these were totally ineffective in activating cPLA2 or LTC4 secretion (Chang & Parekh, 2004; Chang et al. 2007). On the other hand, activation of CRAC channels robustly activated cPLA2 and LTC4 secretion despite raising global Ca2+ to a similar extent (Chang & Parekh, 2004; Chang et al. 2008). We used a nice method developed by Bird & Putney (2005) to raise global Ca2+ after Ca2+ release from the stores. Expulsion of Ca2+ by plasma membrane Ca2+-ATPases is a major Ca2+ clearance mechanism in most cells types. Bird & Putney showed that block of this Ca2+-ATPase with high concentrations of lanthanides rendered cells tight to Ca2+ so that Ca2+ released by thapsigargin remained within the cytosol (Bird & Putney, 2005). Using this approach, we showed that the extent of the global Ca2+ rise following stimulation with thapisigargin and lanthanide in the absence of external Ca2+ was indistinguishable from the rise in response to CRAC channel activation but the latter was severalfold more effective in activating cPLA2 and LTC4 secretion (Chang et al. 2008). Hence local Ca2+ entry through CRAC channels and not global Ca2+ drives this signalling pathway. Consistent with this were the findings that (i) manipulating the electrochemical gradient for Ca2+ influx strongly impacted upon cPLA2 activation with little effect on global Ca2+; (ii) loading cells with the slow Ca2+ chelator EGTA reduced bulk Ca2+ following CRAC channel opening but had little effect on cPLA2 activation, and (iii) although bulk Ca2+ was slightly less after CRAC channel activation in 2 mm Ca2+ and a low concentration of La3+ compared with that seen in 0.5 mm Ca2+, the former condition, which supports larger Ca2+ microdomains, was much more effective in driving LTC4 secretion (Chang et al. 2008).

Sensing local Ca2+ entry

How does a target detect local Ca2+ influx? In general, there are two ways whereby local Ca2+ entry can activate a Ca2+-dependent target. Because the spatial extent of a Ca2+ microdomain can extend tens of nanometres from the channel pore (Neher, 1998b), an effector placed within this range could respond directly to the local Ca2+. Examples include certain Ca2+-activated ion channels and Ca2+-dependent neurotransmitter release (see above), where Ca2+ microdomains near N- and P/Q-type Ca2+ channels bind to synaptotagmin to promote vesicular fusion. Alternatively, a Ca2+ sensor could relay local Ca2+ entry to targets not located within the Ca2+ plume of the open channel. An example here is the coupling of the IQ motif of L-type Ca2+ channels to the ERK pathway (Dolmetsch et al. 2001), which signals to the nucleus. Our work on CRAC channels has identified another way for propagation of local Ca2+ signals into the cytoplasm to activate cPLA2 and leukotriene secretion (Chang et al. 2008). Local Ca2+ entry is detected by the non-receptor tyrosine kinase Syk, which then activates the ERK pathway through recruitment of protein kinase C isoforms α and βII. ERK stimulates cPLA2 by phosphorylating the enzyme on Ser-505 to liberate arachidonic acid (Leslie, 1997). ERK also activates the 5-lipoxygenase enzyme, which metabolizes arachidonic acid to leukotrienes. ERK is therefore an effective and versatile transducer of CRAC channel activity and can bifurcate the local Ca2+ signal into the co-ordinated activation of two important downstream catabolic enzymes. Moreover, dual regulation of these two enzymes ensures that as arachidonic acid is generated, it is rapidly metabolized to LTC4, thereby preventing excessive build-up of this pleiotropic messenger.

Ca2+ channel clustering: overlapping microdomains

Although I have focused exclusively on single Ca2+ channel microdomains, channels may cluster and produce more complex local Ca2+ signals. Now, the local Ca2+ buffers are more rapidly saturated and therefore the spatial extent of the Ca2+ signal increases substantially. In neurons, both a single Ca2+ channel and a cluster of channels drive exocytosis at an active zone. For CRAC channels, the pore-forming subunit Orai1 forms discrete puncta upon store depletion, which would be expected to generate a local Ca2+ concentration much larger than that in response to a single channel. Puncta formation therefore might increase the likelihood of activating targets located further away from the extent of a single microdomain. Interesting modes of regulation could involve varying the number of CRAC channels in a punctum as well as how quickly the puncta form and disaggregate.

Concluding remarks

Local Ca2+ influx provides a mechanism for generating specificity to the promiscuous second messenger Ca2+. Only those effectors located near the channel will be activated by the high ambient Ca2+, thus avoiding a need for a more global and therefore indiscriminate Ca2+ signal. Moreover, selectivity is further increased by targeting effectors to certain Ca2+ channels and not others, as seen with calmodulin tethering to the L-type channel and the tight association between adenylyl cyclases and store-operated channels. Finally, tethering signalling molecules to the Ca2+ channel enables long-range signalling to occur, as occurs with diffusion of calmodulin to the nucleus or following recruitment of protein kinase C to CRAC channels by Syk. Figure 2 summarizes a range of cellular responses activated by local Ca2+ signals near L-type (Fig. 2A) and CRAC channels (Fig. 2B).

Figure 2. The cartoon summarizing the impact of Ca2+ microdomains near L-type (A) and CRAC (B) channels on cell function.

Figure 2

CaM denotes calmodulin, 5LOX 5-lipoxygenase.

Microdomains have other advantages too. Locating the effector near the Ca2+ microdomain greatly increases the speed of response, a mechanism which underlies rapid release of primed neurotransmitter-containing vesicles at the active zone by N- and P/Q-type Ca2+ channels and the shaping of neuronal firing pattern by Ca2+-activated K+ channels colocalized with Ca2+ channels. Microdomains also increase the fidelity of the response; when the Ca2+ channel opens, a juxtaposed effector will generally be exposed to a high local Ca2+. By contrast, the same effector located deeper in the cell may not be exposed to incoming Ca2+ due to Ca2+ buffering and sequestration by intervening mitochondria and ER.

Finally, Ca2+ microdomains are not invariant in amplitude and extent but are controlled by several factors. Clearly, changes in single channel flux through alterations in the electrochemical gradient will have an impact directly on the size of the microdomain and hence its ability to couple to a target. However, one could envisage other, hitherto unexplored, ways of regulating ‘microdomain–effector coupling’. L-type Ca2+ channels show gating mode shifts (Tsien et al. 1986; Armstrong et al. 1991). Mode 0 represents no channel activity despite depolarization, mode 1 reflects brief channel openings of short duration and mode 2 is characterized by continuous channel openings of long duration. Calmodulin-dependent protein kinase-mediated phosphorylation favours stabilization in mode 2 (Dzhura et al. 2000), whereas phosphatases shift channels to mode 0 (Erxleben et al. 2003). The duration of the microdomain as well as the lateral spread would therefore be increased in mode 2 gating. It would be interesting to see whether agonists that stabilize mode 2 enhance the robustness of ‘microdomain–effector coupling’ or recruit new effectors due to the greater spatial spread of the microdomain. Another intriguing possibility arises from the increase in Ca2+ permeability of NMDA receptors upon phosphorylation by protein kinase A, which would increase the size of the Ca2+ microdomain. It is certainly intriguing to consider that coupling between Ca2+ microdomains near NMDA receptors and local targets might be strengthened or weakened depending on the phosphorylation state of the channel. In this way, agonist-specific responses could be elicited despite using the same local Ca2+ signal. Collectively, Ca2+ microdomains provide a simple yet effective mechanism for eliciting specific responses to the pleiotropic and versatile Ca2+ messenger.

Acknowledgments

Work in my laboratory is supported by the Medical Research Council and the British Heart Foundation. I thank Richard Lewis for comments and Daniel Bakowski for help with artwork.

References

  1. Adler EM, Augustine GJ, Duffy SN, Charlton MP. Alien intracellular calcium chelators attenuate neurotransmitter release at the squid giant synapse. J Neurosci. 1991;11:1496–1507. doi: 10.1523/JNEUROSCI.11-06-01496.1991. [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Ahn S, Riccio A, Ginty DD. Spatial considerations for stimulus-dependent transcription in neurons. Annu Rev Neurosci. 2000;21:127–148. doi: 10.1146/annurev.physiol.62.1.803. [DOI] [PubMed] [Google Scholar]
  3. Armstrong DL, Rossier MF, Shcherbatko AD, White RE. Enzymatic gating of voltage-activated calcium channels. Ann NY Acad Sci. 1991;635:26–34. doi: 10.1111/j.1749-6632.1991.tb36478.x. [DOI] [PubMed] [Google Scholar]
  4. Bailey CH, Bartsch D, Kandel ER. Toward a molecular definition of long-term memory storage. Proc Natl Acad Sci U S A. 1996;93:13445–13452. doi: 10.1073/pnas.93.24.13445. [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Bautista DM, Lewis RS. Modulation of plasma membrane calcium-ATPase activity by local calcium microdomains near CRAC channels in human T cells. J Physiol. 2004;556:805–817. doi: 10.1113/jphysiol.2003.060004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  6. Becherer U, Moser T, Stuhmer W, Oheim M. Calcium regulates exocytosis at the level of single vesicles. Nat Neurosci. 2003;6:846–853. doi: 10.1038/nn1087. [DOI] [PubMed] [Google Scholar]
  7. Berkefeld H, Sailer CA, Bildl W, Rhode V, Thumfart JO, Eble S, Klugbauer N, Reisinger E, Bischofberger J, Oliver D, Knaus HG, Schulte U, Fakler B. BKCa-CaV channel complexes mediate rapid and localized calcium-activated potassium signalling. Science. 2006;314:615–620. doi: 10.1126/science.1132915. [DOI] [PubMed] [Google Scholar]
  8. Berridge MJ, Bootman MD, Roderick HL. Calcium signalling: dynamics, homeostasis and remodelling. Nat Rev Mol Cell Biol. 2003;4:517–529. doi: 10.1038/nrm1155. [DOI] [PubMed] [Google Scholar]
  9. Beutner D, Voets T, Neher E, Moser T. Calcium dependence of exocytosis and endocytosis at the cochlea inner hair cell afferent synapse. Neuron. 2004;29:681–690. doi: 10.1016/s0896-6273(01)00243-4. [DOI] [PubMed] [Google Scholar]
  10. Bird GS, Putney JW., Jr Capacitative calcium entry supports calcium oscillations in human embryonic kidney cells. J Physiol. 2005;562:697–706. doi: 10.1113/jphysiol.2004.077289. [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Bollmann J, Sakmann B. Control of synaptic strength and timing by the release-site Ca2+ signal. Nat Neurosci. 2005;8:426–434. doi: 10.1038/nn1417. [DOI] [PubMed] [Google Scholar]
  12. Borst JGG, Sakmann B. Calcium influx and transmitter release in a fast CNS synapse. Nature. 1996;383:431–434. doi: 10.1038/383431a0. [DOI] [PubMed] [Google Scholar]
  13. Chang WC, Di Capite J, Nelson C, Parekh AB. All-or-none activation of CRAC channels by agonist elicits graded responses in populations of mast cells. J Immunol. 2007;179:5255–5263. doi: 10.4049/jimmunol.179.8.5255. [DOI] [PubMed] [Google Scholar]
  14. Chang WC, Di Capite J, Singaravelu K, Nelson C, Halse V, Parekh AB. Local calcium influx through calcium release-activated calcium (CRAC) channels stimulates production of an intracellular messenger and an intercellular pro-inflammatory signal. J Biol Chem. 2008;283:4622–4631. doi: 10.1074/jbc.M705002200. [DOI] [PubMed] [Google Scholar]
  15. Chang WC, Parekh AB. Close functional coupling between CRAC channels, arachidonic acid release and leukotriene secretion. J Biol Chem. 2004;279:29994–29999. doi: 10.1074/jbc.M403969200. [DOI] [PubMed] [Google Scholar]
  16. Chiono M, Mahey R, Tate G, Cooper DMF. Capacitative calcium entry exclusively inhibits cAMP synthesis in C6-2B glioma cells. Evidence that physiologically activated Ca2+ entry regulates Ca2+-inhibitable adenylyl cyclase in non-excitable cells. J Biol Chem. 1995;270:1149–1155. doi: 10.1074/jbc.270.3.1149. [DOI] [PubMed] [Google Scholar]
  17. Church PJ, Stanley EF. Single L-type calcium channel conductance with physiological levels of calcium in chick ciliary ganglion neurons. J Physiol. 1996;496:59–68. doi: 10.1113/jphysiol.1996.sp021665. [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Clapham DE. Calcium signaling. Cell. 2008;131:1047–1058. doi: 10.1016/j.cell.2007.11.028. [DOI] [PubMed] [Google Scholar]
  19. Cull-Candy S, Brickley S, Farrant M. NMDA receptor subunits: diversity, development and disease. Curr Opin Neurobiol. 2001;11:327–335. doi: 10.1016/s0959-4388(00)00215-4. [DOI] [PubMed] [Google Scholar]
  20. Deisseroth K, Bito H, Tsien RW. Signaling from synapse to nucleus: postsynaptic CREB phosphorylation during multiple forms of hippocampal synaptic plasticity. Neuron. 1996;16:89–101. doi: 10.1016/s0896-6273(00)80026-4. [DOI] [PubMed] [Google Scholar]
  21. Deisseroth K, Heist EK, Tsien RW. Translocation of calmodulin to the nucleus supports CREB phosphorylation in hippocampal neurons. Nature. 1998;392:198–202. doi: 10.1038/32448. [DOI] [PubMed] [Google Scholar]
  22. Demuro A, Parker I. Imaging the activity and localization of single voltage-gated Ca channels by total internal reflection fluorescence microscopy. Biophys J. 2004;86:3250–3259. doi: 10.1016/S0006-3495(04)74373-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  23. DiGregorio DA, Vergara JL. Localized detection of action-potential induced presynaptic calcium transients at a Xenopus neuromuscular junction. J Physiol. 1997;505:585–592. doi: 10.1111/j.1469-7793.1997.585ba.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  24. Dolmetsch RE, Pajvani V, Fife K, Spotts JM, Greenberg ME. Signaling to the nucleus by an L-type calcium channel-calmodulin complex through the MAP kinase pathway. Science. 2001;294:333–339. doi: 10.1126/science.1063395. [DOI] [PubMed] [Google Scholar]
  25. Dzhura I, Wu Y, Colbran RJ, Balser JR, Anderson ME. Calmodulin kinase determines calcium-dependent facilitation of L-type calcium channels. Nat Cell Biol. 2000;2:173–177. doi: 10.1038/35004052. [DOI] [PubMed] [Google Scholar]
  26. Erxleben C, Gomez-Alegria C, Darden T, Mori Y, Birnbaumer L, Armstrong DL. Modulation of cardiac Cav1.2 channels by dihydropyridine and phosphatase inhibitor requires Ser-1142 in the domain III pore loop. Proc Natl Acad Sci U S A. 2003;100:2929–2934. doi: 10.1073/pnas.2628046100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. Fagan KA, Mahey R, Cooper DMF. Functional co-localization of transfected Ca2+-stimulable adenylyl cyclases with capacitative Ca2+ entry sites. J Biol Chem. 1996;271:12438–12444. doi: 10.1074/jbc.271.21.12438. [DOI] [PubMed] [Google Scholar]
  28. Fagan KA, Mons N, Cooper DMF. Dependence of the Ca2+-inhibitable adenyl cyclase of C6-2B glioma cells on capacitative Ca2+ entry. J Biol Chem. 1998;273:9297–9305. doi: 10.1074/jbc.273.15.9297. [DOI] [PubMed] [Google Scholar]
  29. Fierro L, Parekh AB. Substantial depletion of the intracellular Ca2+ stores is required for macroscopic activation of the Ca2+ release-activated Ca2+ current in rat basophilic leukaemia cells. J Physiol. 2000;522:247–257. doi: 10.1111/j.1469-7793.2000.t01-1-00247.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  30. Gomez-Ospina N, Tsuruta F, Barreto-Chang O, Hu L, Dolmetsch RE. The C terminus of the L-type voltage-gated calcium channel Cav1.2 encodes a transcription factor. Cell. 2006;127:591–606. doi: 10.1016/j.cell.2006.10.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
  31. Greka A, Navarro B, Oancea E, Duggan A, Clapham DE. TRPC5 is a regulator of hippocampal neurite length and growth cone morphology. Nat Neurosci. 2003;6:837–845. doi: 10.1038/nn1092. [DOI] [PubMed] [Google Scholar]
  32. Gu C, Cooper DMF. Ca2+ Sr2+ and Ba2+ identify distinct regulatory sites on adenylyl cyclase (AC) types VI and VIII and consolidate the apposition of capacitative cation entry channels and Ca2+-sensitive ACs. J Biol Chem. 2000;275:6980–6986. doi: 10.1074/jbc.275.10.6980. [DOI] [PubMed] [Google Scholar]
  33. Hardingham GE, Arnold FJL, Bading H. A calcium microdomain near NMDA receptors: on switch for ERK-dependent synapse-to-nucleus communication. Nat Neurosci. 2001;4:565–566. doi: 10.1038/88380. [DOI] [PubMed] [Google Scholar]
  34. Hardingham GE, Chawla S, Johnson CM, Bading H. Distinct functions of nuclear and cytoplasmic calcium in the control of gene expression. Nature. 1997;385:260–265. doi: 10.1038/385260a0. [DOI] [PubMed] [Google Scholar]
  35. Hardingham GE, Fukunaga Y, Bading H. Extrasynaptic NMDARs oppose synaptic NMDARs by triggering CREB shut-off and cell death pathways. Nat Neurosci. 2002;5:405–414. doi: 10.1038/nn835. [DOI] [PubMed] [Google Scholar]
  36. Heidelberger R, Heinemann C, Neher E, Matthews G. Calcium dependence of the rate of exocytosis in a synaptic terminal. Nature. 1994;371:513–515. doi: 10.1038/371513a0. [DOI] [PubMed] [Google Scholar]
  37. Hoth M, Penner R. Depletion of intracellular calcium stores activates a calcium current in mast cells. Nature. 1992;355:353–356. doi: 10.1038/355353a0. [DOI] [PubMed] [Google Scholar]
  38. Hu S-C, Chrivia J, Ghosh A. Regulation of CBP-mediated transcription by neuronal calcium signalling. Neuron. 1999;22:799–808. doi: 10.1016/s0896-6273(00)80738-2. [DOI] [PubMed] [Google Scholar]
  39. Lansman JB, Hess P, Tsien RW. Blockade of current through single calcium channels by Cd2+, Mg2+, and Ca2+. Voltage and concentration dependence of calcium entry into the pore. J Gen Physiol. 1986;88:321–347. doi: 10.1085/jgp.88.3.321. [DOI] [PMC free article] [PubMed] [Google Scholar]
  40. Lattanzio FA, Jr, Bartschat DK. The effect of pH on rate constants, ion selectivity and thermodynamic properties of fluorescent calcium and magnesium indicators. Biochem Biophys Res Commun. 1991;177:184–191. doi: 10.1016/0006-291x(91)91966-g. [DOI] [PubMed] [Google Scholar]
  41. Lee A, Wong ST, Gallagher D, Li B, Storm DR, Scheuer T, Catterall WA. Ca2+/calmodulin binds to and modulates P/Q-type calcium channels. Nature. 1999;399:155–159. doi: 10.1038/20194. [DOI] [PubMed] [Google Scholar]
  42. Lemonnier L, Prevarskaya N, Shuba Y, Vanden-Abeele F, Nilius B, Mazurier J, Skryma R. Ca2+ modulation of volume-regulated anion channels: evidence for colocalization with store-operated channels. FASEB J. 2002;16:222–224. doi: 10.1096/fj.01-0383fje. [DOI] [PubMed] [Google Scholar]
  43. Lemonnier L, Trebak M, Lievremont JP, Bird GS, Putney JW., Jr Protection of TRPC7 cation channels from calcium inhibition by closely associated SERCA pumps. FASEB J. 2006;20:503–505. doi: 10.1096/fj.05-4714fje. [DOI] [PubMed] [Google Scholar]
  44. Leslie CC. Properties and regulation of cytosolic phospholipase A2. J Biol Chem. 1997;272:16709–16712. doi: 10.1074/jbc.272.27.16709. [DOI] [PubMed] [Google Scholar]
  45. Lewis RS. The molecular choreography of a store-operated calcium channel. Nature. 2007;446:284–287. doi: 10.1038/nature05637. [DOI] [PubMed] [Google Scholar]
  46. Li Y, Jia YC, Cui K, Li N, Zheng ZY, Wang YZ, Yuan XB. Essential role of TRPC channels in the guidance of nerve growth cones by brain-derived neurotrophic factor. Nature. 2005;434:894–898. doi: 10.1038/nature03477. [DOI] [PubMed] [Google Scholar]
  47. Lin S, Fagan KA, Li K-X, Shaul WPW, Cooper DMF, Rodman DM. Sustained endothelial nitric oxide-synthase activation requires capacitative Ca2+ entry. J Biol Chem. 2000;275:17979–17985. doi: 10.1074/jbc.275.24.17979. [DOI] [PubMed] [Google Scholar]
  48. Llinas R, Sugimori M, Silver RB. Microdomains of high calcium concentration in a presynaptic terminal. Science. 1992;256:677–679. doi: 10.1126/science.1350109. [DOI] [PubMed] [Google Scholar]
  49. Martin AC, Cooper DMF. Capacitative and 1-oleyl-2-acetyl-sn-glycerol-activated Ca2+ entry distinguished using adenyl cyclase. Mol Pharmacol. 2006;70:769–777. doi: 10.1124/mol.106.025213. [DOI] [PubMed] [Google Scholar]
  50. Meinrenken C, Borst JGG, Sakmann B. Calcium secretion coupling at the calyx of Held governed by nonuniform channel-vesicle topography. J Neurosci. 2002;22:1648–1667. doi: 10.1523/JNEUROSCI.22-05-01648.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  51. Mermelstein PG, Bito H, Deisseroth K, Tsien RW. Critical dependence of cAMP response element-binding protein phosphorylation on L-type calcium channels supports a selective response to EPSPs in preference to action potentials. J Neurosci. 2000;20:266–273. doi: 10.1523/JNEUROSCI.20-01-00266.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  52. Mogami H, Gardner J, Gerasimenko OV, Camello P, Petersen OH, Tepikin AV. Calcium binding capacity of the cytosol and endoplasmic reticulum of mouse pancreatic acinar cells. J Physiol. 1999;518:463–467. doi: 10.1111/j.1469-7793.1999.0463p.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  53. Mogami H, Nakano K, Tepikin AV, Petersen OH. Ca2+ flow via tunnels in polarized cells: recharging of apical Ca2+ stores by focal Ca2+ entry through basal membrane patch. Cell. 1997;88:49–55. doi: 10.1016/s0092-8674(00)81857-7. [DOI] [PubMed] [Google Scholar]
  54. Neher E. The use of Fura 2 for estimating Ca2+ buffers and Ca2+ fluxes. Neuropharmacology. 1995;34:1423–1442. doi: 10.1016/0028-3908(95)00144-u. [DOI] [PubMed] [Google Scholar]
  55. Neher E. Usefulness and limitations of linear approximations to the understanding of Ca2+ signals. Cell Calcium. 1998a;24:345–357. doi: 10.1016/s0143-4160(98)90058-6. [DOI] [PubMed] [Google Scholar]
  56. Neher E. Vesicle pools and Ca2+ microdomains: new tools for understanding their roles in neurotransmitter release. Neuron. 1998b;20:389–399. doi: 10.1016/s0896-6273(00)80983-6. [DOI] [PubMed] [Google Scholar]
  57. Nilius B, Owsianik G, Voets T, Peters JA. Transient receptor potential cation channels in disease. Physiol Rev. 2007;87:165–217. doi: 10.1152/physrev.00021.2006. [DOI] [PubMed] [Google Scholar]
  58. Parekh AB. Functional consequences of activating store-operated CRAC channels. Cell Calcium. 2007;42:111–121. doi: 10.1016/j.ceca.2007.02.012. [DOI] [PubMed] [Google Scholar]
  59. Parekh AB, Putney JW., Jr Store-operated calcium channels. Physiol Rev. 2005;85:757–810. doi: 10.1152/physrev.00057.2003. [DOI] [PubMed] [Google Scholar]
  60. Petersen OH. Calcium signalling and calcium-activated ion channels in exocrine acinar cells. Cell Calcium. 2005;38:171–200. doi: 10.1016/j.ceca.2005.06.024. [DOI] [PubMed] [Google Scholar]
  61. Petersen OH, Tepikin A, Park MK. The endoplasmic reticulum: one contiuous or several separate Ca2+ stores. Trends Neurosci. 2001;24:271–276. doi: 10.1016/s0166-2236(00)01787-2. [DOI] [PubMed] [Google Scholar]
  62. Peters-Golden M, Gleason MM, Togias A. Cysteinyl leukotrienes: multi-functional mediators in allergic rhinitis. Clin Exp Allergy. 2006;36:689–703. doi: 10.1111/j.1365-2222.2006.02498.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  63. Prakriya M, Solaro CR, Lingle CJ. [Ca2+]i elevations detected by BK channels during Ca2+ influx and muscarine-mediated reelase of Ca2+ from intracellular Ca2+ stores in rat chromaffin cells. J Neurosci. 1996;16:4344–4359. doi: 10.1523/JNEUROSCI.16-14-04344.1996. [DOI] [PMC free article] [PubMed] [Google Scholar]
  64. Richardson A, Taylor CW. Effects of calcium chelators on purified inositol1,4,5-trisphosphate (InsP3) receptors and InsP3 stimulated calcium mobliization. J Biol Chem. 1993;268:11528–11533. [PubMed] [Google Scholar]
  65. Rizzuto R, Pozzan T. Microdomains of intracellular calcium: molecular determinants and functional consequences. Physiol Rev. 2006;86:369–408. doi: 10.1152/physrev.00004.2005. [DOI] [PubMed] [Google Scholar]
  66. Roberts WM, Jacobs RA, Hudspeth AJ. Colocalization of ion channels involved in frequency selectivity and synaptic transmission of presynaptic active zones of hair cells. J Neurosci. 1990;10:3664–3684. doi: 10.1523/JNEUROSCI.10-11-03664.1990. [DOI] [PMC free article] [PubMed] [Google Scholar]
  67. Robitaille E, Adler EM, Charlton MP. Strategic location of calcium channels and transmitter release sites of frog neuromuscular synapses. Neuron. 1990;5:773–779. doi: 10.1016/0896-6273(90)90336-e. [DOI] [PubMed] [Google Scholar]
  68. Rozov A, Burnashev N, Sakmann B, Neher E. Transmitter release modulation by intracellular Ca2+ buffers in facilitating and depressing nerve terminals of pyramidal cells in layer 2/3 of the rat neocortex indicates a target cell-specific difference in presynaptic dynamics. J Physiol. 2001;531:807–826. doi: 10.1111/j.1469-7793.2001.0807h.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  69. Schneggenburger R, Neher E. Intracellular calcium dependence of transmitter release rates at a fast central synapse. Nature. 2000;406:889–893. doi: 10.1038/35022702. [DOI] [PubMed] [Google Scholar]
  70. Shaywitz AJ, Greenberg ME. A stimulus-induced transcription factor activated by a diverse array of extracellular signals. Annu Rev Biochem. 1999;68:821–861. doi: 10.1146/annurev.biochem.68.1.821. [DOI] [PubMed] [Google Scholar]
  71. Shuttleworth TJ, Thompson JL. Discriminating between capacitative and arachidonate-activated Ca2+ entry pathways in HEK293 cells. J Biol Chem. 1999;274:31174–31178. doi: 10.1074/jbc.274.44.31174. [DOI] [PubMed] [Google Scholar]
  72. Skeberdis VA, Chevaleyre V, Lau CG, Goldberg JH, Pettit DL, Suadicani SO, Lin Y, Bennett MVL, Yuste R, Castillo PE, Zukin RS. Protein kinase A regulates calcium permeability of NMDA receptors. Nat Neurosci. 2006;9:501–510. doi: 10.1038/nn1664. [DOI] [PubMed] [Google Scholar]
  73. Spitzer NC. Electrical activity in early neuronal development. Nature. 2006;444:707–712. doi: 10.1038/nature05300. [DOI] [PubMed] [Google Scholar]
  74. Sutton KG, McRory JE, Guthrie H, Murphy TH, Snutch TP. P/Q-type calcium channels mediate the activity-dependent feedback of syntaxin-1A. Nature. 1999;401:800–804. doi: 10.1038/44586. [DOI] [PubMed] [Google Scholar]
  75. Takasu MA, Dalva MB, Zigmond RE, Greenberg ME. Modulation of NMDA receptor-dependent calcium influx and gene expression through EphB receptors. Science. 2002;295:491–495. doi: 10.1126/science.1065983. [DOI] [PubMed] [Google Scholar]
  76. Tsien RW, Bean BP, Hess P, Lansman JB, Nilius B, Nowycky MC. Mechanisms of calcium channel modulation by β-adrenergic agonists and dihydropyridine calcium agonists. J Mol Cell Cardiol. 1986;18:691–710. doi: 10.1016/s0022-2828(86)80941-5. [DOI] [PubMed] [Google Scholar]
  77. Wang GX, Poo MM. Requirement of TRPC channels in netrin-1-induced chemotropic turning of nerve growth cones. Nature. 2005;434:835–838. doi: 10.1038/nature03478. [DOI] [PubMed] [Google Scholar]
  78. Weick JP, Groth RD, Isaksen AL, Mermelstein PG. Interactions with PDZ proteins are required for L-type calcium channels to activate cAMP element-binding protein-dependent gene expression. J Neurosci. 2003;23:3446–3456. doi: 10.1523/JNEUROSCI.23-08-03446.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  79. Willoughby D, Cooper DMF. Organization and Ca2+ regulation of adenylyl cyclases in cAMP microdomains. Physiol Rev. 2007;87:965–1010. doi: 10.1152/physrev.00049.2006. [DOI] [PubMed] [Google Scholar]
  80. Zenisek D, Davila V, Wan L, Almers W. Imaging calcium entry sites and ribbon structures in two presynaptic cells. J Neurosci. 2003;23:2538–2548. doi: 10.1523/JNEUROSCI.23-07-02538.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  81. Zheng JQ. Turning of nerve growth cones induced by localized increases in intracellular calcium ions. Nature. 2000;403:89–93. doi: 10.1038/47501. [DOI] [PubMed] [Google Scholar]
  82. Zuehlke 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]
  83. Zweifach A, Lewis RS. Mitogen-regulated Ca2+ current of T lymphocytes is activated by depletion of intracellular Ca2+ stores. Proc Natl Acad Sci U S A. 1993;90:6295–6299. doi: 10.1073/pnas.90.13.6295. [DOI] [PMC free article] [PubMed] [Google Scholar]
  84. Zweifach A, Lewis RS. Rapid inactivation of depletion-activated calcium current (ICRAC) due to local calcium feedback. J Gen Physiol. 1995;105:209–226. doi: 10.1085/jgp.105.2.209. [DOI] [PMC free article] [PubMed] [Google Scholar]

Articles from The Journal of Physiology are provided here courtesy of The Physiological Society

RESOURCES