Abstract
Modulation of cytoplasmic free Ca2+ concentration ([Ca2+]i) by receptor-mediated generation of inositol 1,4,5-trisphosphate (InsP3) and activation of its receptor (InsP3R), a Ca2+-release channel in the endoplasmic reticulum, is a ubiquitous signalling mechanism. A fundamental aspect of InsP3-mediated signalling is the graded release of Ca2+ in response to incremental levels of stimuli. Ca2+ release has a transient fast phase, whose rate is proportional to [InsP3], followed by a much slower one even in constant [InsP3]. Many schemes have been proposed to account for quantal Ca2+ release, including the presence of heterogeneous channels and Ca2+ stores with various mechanisms of release termination. Here, we demonstrate that mechanisms intrinsic to the single InsP3R channel can account for quantal Ca2+ release. Patch-clamp electrophysiology of isolated insect Sf9 cell nuclei revealed a consistent and high probability of detecting functional endogenous InsP3R channels, enabling InsP3-induced channel inactivation to be identified as an inevitable consequence of activation, and allowing the average number of activated channels in the membrane patch (NA) to be accurately quantified. InsP3-activated channels invariably inactivated, with average duration of channel activity reduced by high [Ca2+]i and suboptimal [InsP3]. Unexpectedly, NA was found to be a graded function of both [Ca2+]i and [InsP3]. A qualitative model involving Ca2+-induced InsP3R sequestration and inactivation can account for these observations. These results suggest that apparent heterogeneous ligand sensitivity can be generated in a homogeneous population of InsP3R channels, providing a mechanism for graded Ca2+ release that is intrinsic to the InsP3R Ca2+ release channel itself.
Modulation of cytoplasmic free Ca2+ concentration ([Ca2+]i) is a ubiquitous signalling mechanism that regulates many physiological processes, ranging from gene expression to apoptosis. Receptor-mediated hydrolysis of phosphatidylinositol 4,5-bisphosphate generates inositol 1,4,5-trisphosphate (InsP3), which binds to its receptor (InsP3R), a ligand-gated Ca2+-release channel in the endoplasmic reticulum (ER). Analyses of InsP3-mediated [Ca2+]i signals in single cells have revealed them to be extremely complex temporally and spatially, providing different signals to discrete parts of the cell (Berridge, 1993).
A fundamental aspect of InsP3R-mediated intracellular signalling is the phenomenon of ‘quantal release’, defined here as the ability of cells to have graded release of Ca2+ from intracellular stores in response to incremental levels of extracellular agonist or cytoplasmic InsP3 concentrations ([InsP3]) (Muallem et al. 1989; Meyer & Stryer, 1990) (reviewed in Bootman, 1994; Missiaen et al. 1994; Parys et al. 1996; Taylor, 1998). Ca2+ release in response to InsP3 has a transient fast phase, whose rate is proportional to [InsP3], followed by a much slower one even in constant [InsP3]. Consequently, sustained exposure to submaximal levels of agonists, even over extensive periods, only mobilizes a fraction of total releasable Ca2+ in a cell. Quantal Ca2+ release has been observed in many cell types under a variety of experimental conditions, including cells with the plasma membrane permeabilized, or with ATP depleted, or with [Ca2+]i kept constant, and in response to poorly metabolizable InsP3 analogues (Bootman, 1994; Missiaen et al. 1994; Parys et al. 1996; Taylor, 1998). Quantal Ca2+ release is surprising because it might have been expected that all InsP3R channels should become activated in response to any agonist concentration, releasing all of the InsP3-sensitive Ca2+ stores, albeit at different rates depending on the agonist concentration. Despite considerable investigation, there is no consensus regarding the mechanisms that either tune the initial rate of release of Ca2+ to [InsP3], or subsequently significantly slow or terminate release. Several mechanisms involving either InsP3R or Ca2+ store heterogeneity have been invoked to account for graded transient Ca2+ release, including the presence of discrete Ca2+ stores with different densities of InsP3R or sensitivities to [InsP3], or the presence of heterogeneous InsP3R channels in a continuous store (different channel isoforms with alternatively spliced variants and variable post-translational modifications have been proposed) with different proposed mechanisms of release termination, including regulation of InsP3R activity by ER luminal [Ca2+] or by desensitization or inactivation processes (Bootman, 1994; Missiaen et al. 1994; Parys et al. 1996; Taylor, 1998). Nevertheless, the mechanisms involved in graded transient Ca2+ release have remained controversial because experimental evidence has been obtained that has disputed every proposed scheme (Bootman, 1994; Missiaen et al. 1994; Parys et al. 1996; Taylor, 1998).
High-resolution electrophysiological and imaging studies have provided more recent insights into the mechanisms of graded release. First, optical imaging of [Ca2+] indicator dyes has revealed the presence of highly localized Ca2+ release events whose numbers and frequency are graded with stimulus intensity (Parker & Ivorra, 1990b; Lechleiter et al. 1991; Bootman & Berridge, 1996; Horne & Meyer, 1997; Sun et al. 1998; Thomas et al. 1998). Second, recordings of single InsP3R channels have suggested that feedback inhibition of channel gating by high [Ca2+]i, tuned by [InsP3], can grade channel activity with [InsP3] (Mak et al. 1998, 2001). Together, these observations provide a mechanism to grade Ca2+ release that is based on ligand-dependent activity of independent release sites. Thus, in the presence of low [InsP3] during weak agonist stimulation, sporadic random openings of InsP3R channels (Mak et al. 1998, 2001, 2003a) generate small localized [Ca2+]i elevations (termed ‘blips’) (Parker & Yao, 1996; Bootman et al. 1997). As [InsP3] increases, sensitivity of the InsP3R to Ca2+ inhibition is reduced (Mak et al. 1998, 2001), enabling Ca2+-induced Ca2+ release (CICR) (Iino, 1990) to coordinate openings of neighbouring InsP3Rs in a local cluster (Yao et al. 1995; Horne & Meyer, 1997; Mak & Foskett, 1997; Thomas et al. 1998) to generate larger Ca2+‘puffs’ (Parker & Yao, 1995; Yao et al. 1995; Sun et al. 1998). Puffs remain localized when the level of released Ca2+ is insufficient to allow Ca2+ diffusing from puffs to activate InsP3Rs beyond the local cluster (Parker & Yao, 1996), whereas even greater Ca2+ release at higher [InsP3] can enable Ca2+ to diffuse far enough to activate channels in other clusters. This progressive recruitment of release sites by even higher [InsP3] may then account for transitions from local to global Ca2+ signals (Bootman & Berridge, 1995; Berridge, 1997). Although this model can account for graded Ca2+ release, it is not known if InsP3-tuning of high [Ca2+] inhibition is the primary mechanism used, or whether additional mechanisms also contribute to InsP3R channel recruitment. Furthermore, this model does not account for observed apparent heterogeneity of InsP3 sensitivity among release sites throughout different regions of the cytoplasm (Bootman & Berridge, 1996; Parker et al. 1996; Thorn et al. 1996), nor does it provide insights into the mechanisms that underlie time-dependent termination of Ca2+ release.
Here, we demonstrate that mechanisms intrinsic to the single InsP3R channel itself can account for quantal Ca2+ release. Patch-clamp electrophysiology of nuclei isolated from insect Sf9 cells yielded a consistent and high probability of detecting apparently homogenous endogenous InsP3R channels, which has enabled two novel observations. First, by accurate determinations of channel activity durations, we have established InsP3-induced InsP3R inactivation as an inevitable process that terminates single channel activity, with the rate of inactivation regulated by [InsP3] and [Ca2+]i. We established this ligand-dependent process as true inactivation by demonstrating its reversibility. Second, we have been able to accurately quantify the mean number of activated channels in a typical membrane patch under precisely controlled ligand conditions. Whereas it was anticipated that all channels in a homogeneous population would always become activated, even in suboptimal ligand conditions, albeit to lower levels of activity, we found instead that the number of channels activated is a graded function of both [Ca2+]i and [InsP3]. A qualitative model can account for graded channel recruitment by suggesting that apparent heterogeneous ligand sensitivities can be generated in a homogeneous population of InsP3R channels. This model also accounts for the ligand concentration dependence of the rate of fateful InsP3-induced channel inactivation. Thus, our results suggest that quantal Ca2+ release can be generated by mechanisms that are intrinsic to the InsP3R Ca2+-release channel itself.
Methods
Spodoptera frugiperda (Sf9) cells (Invitrogen) were grown and maintained in SF-900II serum-free media (Gibco) in suspension culture according to the manufacturer's protocols. For maximum detection of functional InsP3R channels in patch-clamp experiments, each freshly thawed batch of cells was passaged 3–4 times before being used for electrophysiology. Cells were propagated for up to 7–8 weeks in culture before a new batch was thawed and expanded. Cells were moved from suspension culture to a T-25 flask, allowed to attach to the bottom of the flask for 1 h, and then washed twice with Ca2+- and Mg2+-free PBS. An ice-cold nuclear isolation solution, containing (mm): 140 KCl, 250 sucrose, 1.5 β-mercaptoethanol, 10 Tris-HCl (pH adjusted to 7.4), with complete protease inhibitor cocktail (Roche Molecular Biochemicals, Indianapolis, IN, USA) and 0.05 mm phenylmethylsulphonylfluoride (PMSF), was added to the flask and the cells were detached by gentle scraping. Homogenization of 1–2 ml of the mixture was performed using 2–4 strokes of the pestle in an ice-cold Dounce homogenizer. A 20–30 μl volume of the homogenized mixture was added to 1 ml standard bath solution (mm: 140 KCl, 10 Hepes, 0.5 BAPTA, pH 7.3, and free [Ca2+] adjusted to ∼300 nm) in an experimental chamber on the stage of an inverted microscope. Isolated nuclei are ∼10 μm in diameter, and were distinguished from intact cells based on their unique morphology (Fig. 1B). Fresh homogenizations were performed every 2 h (Mak et al. 2005). The standard pipette solution contained (mm): 140 KCl, 0.5 Na2ATP, 10 Hepes, pH 7.3, and various [Ca2+] and [InsP3], as indicated. All solutions were carefully buffered to desired free [Ca2+] (Mak et al. 1998), confirmed by fluorometry.
Figure 1. Single-channel recording of Sf9-InsP3R in isolated nuclei.
A, typical multiple Sf9-InsP3R channel current trace. Five channels, as indicated by open channel current levels (arrows), were activated at beginning of record. Lowest arrow denotes zero-current level. For this particular record, channel activity was recorded for 210 s from the beginning of the record to the last channel closing event. Channels inactivated sequentially until all were inactivated at the end of the record. Data were recorded at +20 mV holding potential in 10 μm InsP3 and 1 μm Ca2+i. An InsP3R channel was in a subconductance level in the shaded region. B, identification of nuclei for patch-clamp electrophysiology. In the cell homogenate transferred to the recording chamber, isolated nucleus (arrow) and intact (1) and partially damaged (2) cells were present and morphologically distinguishable. Scale bar, 10 μm. C, single InsP3R channel current recording during voltage ramp (from −12.5 to +12.5 mV). Open circles and dashed line indicate linear fits of open and closed channel current–voltage (I–V) relations, respectively. The record was obtained in standard bath and pipette solutions, with 10 μm InsP3 and 1 μm Ca2+i in the pipette solution. D, analysis of number of channels observed in 75 membrane patches with 10 μm InsP3 and 1 μm Ca2+i. The open bars show observed distribution with average 2.05 channels per patch. The shaded histogram is the theoretical Poisson distribution with same number of membrane patches and mean number of channels per patch.
Data were acquired as described (Mak et al. 1998). Segments of current traces exhibiting one or two InsP3R channels were used for open probability (Po) determinations (Mak & Foskett, 1997), and single-channel traces were used for dwell-time analyses by QuB software (Qin et al. 2000). To determine relative ionic permeabilities, patches were excised into normal bath solution, and then exposed to a solution containing (mm): 110 N-methyl-d-glucamine (NMDG) chloride, 30 KCl, 0.192 CaCl2, 0.5 BAPTA, 10 Hepes, pH 7.3 (for PK:PCl determinations, where P is permeability); 140 KCl, 10 CaCl2, 10 Hepes, pH 7.3 (for PK:PCa determinations); 140 KCl, 10 MgCl2, 0.192 CaCl2, 10 Hepes, 0.5 BAPTA, pH 7.3 (for PK:PMg determinations). Independent polynomial fits (Igor Pro WaveMetrics) of channel open and closed current levels during voltage ramps (10–20 sweeps at 100 mV s−1, from −70 to +30 mV (PK:PCl) or from −50 to +50 mV (PK:Pdivalent) were used to determine reversal potentials, which were corrected for liquid junction potentials. Permeability ratios were calculated using the Goldman-Hodgkin-Katz (GHK) equation for either two or three ion species.
Particular consideration was given to the accurate determination of the number of active channels in nuclear membrane patches (NA) from the experimental current records. For simplicity, we consider a membrane patch containing n active, identical and independent channels exhibiting Markovian stochastic kinetics with open probability Po and a single open channel kinetic state with mean open duration of to. With these assumptions, events when all n channels open simultaneously follow Markovian kinetics with a probability (Po)n and mean duration of to/n. If T is the minimum duration of an open event that is detectable after filtering in the experimental system, then average interval σn between two successive detectable events when all n channels were open is
. Therefore, the number of active channels in a membrane patch (N) can be assumed, with high level of confidence (P < 0.01), to be the maximum number of open channel current levels observed if the channel record lasted longer than 5(σN+1). In our patch-clamp set-up, T was empirically determined to be 0.2 ms using test pulses of variable duration (Mak et al. 2001). The to value of the Sf9 InsP3R channels is ∼30 ms over the range of experimental conditions used. In experiments performed using conditions that generated the lowest channel Po (0.09 at [InsP3]= 33 nm and [Ca2+]i= 7.5 μm), at most one active channel was involved, and the current records analysed all lasted longer than 5(σ2). In experiments using conditions that produced higher channel Po, multi-channel current records were more common, but the duration of each current record analysed exceeded 5(σN+1), where N is the number of maximum open channel current levels detected in that record. Thus, the uncertainty in the determination of the number of active channels in a membrane patch was insignificant in the studies described here.
We observe that gating activities of all InsP3R channels inevitably terminate, i.e. InsP3R channel activities observed in nuclear patch-clamp experiments have finite durations. To quantify the activity duration for an InsP3R channel in an experiment, we assume that all channels observed in one experiment inactivated identically, independently, following simple Markovian kinetics with a single time constant (Ta). Then, for an experiment with N active channels and channel activity lasting TN from the beginning of the experiment to the last channel closing observed in the channel current record,
.
Besides the durations of observation of channel activities, another factor affecting the accurate determination of the number of active channels in nuclear membrane patches is the finite time ts between the initial exposure of the InsP3R channels to the ligand conditions in the pipette solution as the tip of the patch-clamp pipette made contact with the outer nuclear envelope, and the quality of seal between the isolated membrane patch and the patch pipette tip becoming good enough (seal resistance >200 MΩ) for single-channel current recording. Because of the inactivation of the InsP3R channels in every experiment, a fraction of all InsP3R channels stimulated by ligand conditions in the patch pipette were not observed because they underwent inactivation during ts. For isolated Sf9 nuclei, mean ts was 6.0 s (standard deviation 2.8 s). With the same assumptions used in Ta evaluation, the actual number of InsP3R channels activated in an experiment NA was estimated as Nobs exp(ts/Ta), where Nobs is the number of InsP3R channels observed in the current record. For most ligand conditions ts≪Ta so that the values of NA are not significantly different from those of Nobs, except in 10 μm InsP3 and 89 μm Ca2+i, when Ta is so short (9.1 s) that NA cannot be confidently estimated from Nobs.
All data points shown in all graphs are means of at least four experiments performed at the same [InsP3] and [Ca2+]i. Error bars indicate s.e.m. The s.e.m. for NA under one set of ligand conditions was calculated by taking into account the effects of Nobs, Ta and ts.
Results
Single-channel recording of native InsP3R in Sf9 cell nuclear membranes
Single InsP3R channels have been recorded in native ER membranes by patch-clamping nuclei isolated from Xenopus oocytes (Mak & Foskett, 1997) and COS-7 cells (Boehning et al. 2001a). This technique is rationalized by the continuity of the ER with the outer membrane of the nuclear envelope and the role of the nuclear envelope as a Ca2+ store similar to ER. Shortcomings of the previously used Xenopus and COS-7 cell systems include lack of consistency in detecting active InsP3R channels and relatively rapid (t½ <30 s) apparent channel inactivation following InsP3-induced activation (Mak & Foskett, 1997; Mak et al. 2000; Boehning et al. 2001a). To discover a better system for studying single InsP3R channels, we examined insect Sf9 cells. Gigaohm seals were readily achieved (>80% success rate) on carefully selected isolated nuclei (Fig. 1B). With 10 μm InsP3 and 1 μm Ca2+ in the pipette solution, conditions that maximize activity of rat and Xenopus InsP3R channels in nuclear membrane patches (Mak et al. 1998, 2001; Boehning et al. 2001a), Sf9 InsP3R channels were consistently detected in 60–80% of patches obtained from nuclei isolated from different batches of cells. Channel activity was evident upon gigaseal formation, but the number of active channels decreased during continuous recording (Fig. 1A) due to abrupt activity termination. The channels had linear current–voltage relationship with slope conductance of 477 ± 3 pS (n = 10) (Fig. 1C), larger than the 360 pS of vertebrate channels (Mak & Foskett, 1998). Channels varied somewhat in their conductance, which sometimes fluctuated during a continuous recording (Mak & Foskett, 1998) (s.d. 12 pS), with occasional openings to subconductance levels that accounted for <2% of all open durations (Fig. 1A). The permeability sequence, determined from reversal potential measurements of currents in asymmetrical solutions, PCa:PMg:PK:PCl was 10:6.8:1:0.22. No channels were observed in the absence of InsP3 (0/13 patches) or presence of 10 μm InsP3 and the competitive inhibitor heparin (100 μg ml−1) (0/14 patches), whereas with 10 μm InsP3 they were detected in 5/6 patches obtained from the same batch of cells. Thus, the channels observed were confirmed to be endogenous Sf9 InsP3R channels.
Distribution of InsP3R channels on Sf9 nuclei
A significant proportion (41/53) of active patches in optimal conditions (10 μm InsP3, 1 μm Ca2+) contained multiple InsP3R channels, as identified by two or more equally spaced current levels, each corresponding to the predominant ∼480 pS state (Fig. 1A). Comparing the number of channels observed in 75 patches with the theoretical Poisson distribution revealed a significant disparity between the two (Fig. 1D), suggesting that the channels are not evenly distributed in the outer nuclear membrane. Thus, Sf9-InsP3R channels may be organized in clusters, a conclusion similar to that reached in an analogous analysis of Xenopus oocyte InsP3R in outer nuclear membrane (Mak & Foskett, 1997) and in Ca2+-imaging studies of several cell types (Yao et al. 1995; Horne & Meyer, 1997; Thomas et al. 1998; Bootman et al. 2001).
[Ca2+] and [InsP3] dependencies of Sf9-InsP3R channel gating kinetics
InsP3R channel gating is under complex allosteric regulation by both InsP3 and [Ca2+]i. Activity of Sf9-InsP3R is biphasically regulated by [Ca2+]i with maximum channel Po exhibited over a broad range of [Ca2+]i in the presence of saturating [InsP3] (Fig. 2A–C). The [Ca2+]i dependence of Po in the presence of 10 μm InsP3 can be fitted with a biphasic Hill equation (Fig. 2B) (Mak et al. 1998):
Figure 2. Ligand regulation of the Sf9-InsP3R channel activities.
A, representative current traces at various [Ca2+]i with channels stimulated by saturating (10 μm) and subsaturating (33 nm) [InsP3], as indicated. B, dependence of InsP3R channel Po on [Ca2+]i and [InsP3], with curves representing empirical fits to the observed channel Po using the biphasic Hill equation (eqn (1)) with parameters shown in Table 1. Inset, [InsP3] dependence of Kinh derived from biphasic Hill equation fits to data, fitted with a simple activation Hill equation (Mak et al. 1998), with parameters: apparent half-maximal InsP3 concentration (KIP3) and Hill coefficient (HIP3) as tabulated. C, dependence of InsP3R channel Po on [Ca2+]i and [InsP3], with curves representing the theoretical fits to the observed channel Po using the Monod-Wyman-Changeux (MWC)-based allosteric model described in Mak et al. (2003a). D, dependence on [Ca2+]i and [InsP3] of mean open channel duration (to) and mean closed channel duration (tc). Data points are obtained under saturating and subsaturating [InsP3] as tabulated.
![]() |
1 |
with the Hill equation parameters – maximum Po (Pmax), half-maximal activating [Ca2+]i (Kact), activation Hill coefficient (Hact), half-maximal inhibitory [Ca2+]i (Kinh), and inhibition Hill coefficient (Hinh) – tabulated in Table 1. As [InsP3] was reduced, the observed maximum Po was reduced and the channel activity was more sensitive to high-[Ca2+]i inhibition (Fig. 2A and B). The Poversus[Ca2+]i dependencies in subsaturating [InsP3] were well-fitted using the same biphasic Hill equations with Kinh dramatically sensitive to [InsP3]: Kinh decreased by nearly three orders of magnitude from 26 μm at 10 μm InsP3 to 82 nm at 10 nm InsP3 (Fig. 2B and Table 1). In contrast, the Ca2+ activation properties of Sf9-InsP3R were not significantly affected by changes in [InsP3] (Table 1). Thus, the main effect of InsP3 on Sf9-InsP3R channel activity was to reduce the sensitivity of the channel to inhibition by [Ca2+]i. The half-maximal activating concentration of InsP3, determined from a monophasic Hill equation fit of the InsP3 dependence of Kinh, was 390 nm with Hill coefficient of 1.7 (Fig. 2B). Kinetically, [Ca2+]i and [InsP3] regulate the Sf9 InsP3R channel Po mainly by affecting the mean closed channel duration tc, with relatively minor effects on mean open channel duration to (Fig. 2D). Dependence of Sf9-InsP3R channel gating kinetics (Po, to and tc) on [Ca2+]i and [InsP3] are remarkably similar to that observed in its vertebrate homologs (Mak et al. 1998, 2001; Boehning et al. 2001a). Notably, the complex ligand regulation of the Sf9 InsP3R channel gating kinetics can be well described by an allosteric Monod-Wyman-Changeux (MWC)-based model that was previously developed to account for ligand regulation of the vertebrate InsP3Rs (Fig. 2C) (Mak et al. 2003a).
Table 1.
Hill equation parameters for [Ca2+] dependence of insect Sf9 InsP3R at different [InsP3]
| [InsP3] | Pmaxa | Kact (nm)b | Hactb | Kinh (μm)b | Hinhb |
|---|---|---|---|---|---|
| 10 μm | 0.75 | 200 | 1.4 | 30 | 1.1 |
| 100 nm | 0.75 | 280 | 1.4 | 2.2 | 1.2 |
| 33 nm | 0.75 | 300 | 1.4 | 0.38 | 1.0 |
| 10 nm | 0.75 | 250 | 1.4 | 0.08 | 1.3 |
Pmax (maximum open probability) was constant for all [InsP3] because according to the Monod-Wyman-Changeux-based model for the InsP3R, the value of Pmax is determined by ligand-independent transitions between various channel configurations, and therefore exhibits no ligand dependence.
The values of Kact (half-maximal activating [Ca2+]i), Kinh (half-maximal inhibitory [Ca2+]i) and Hinh (inhibition Hill coefficient) were determined by fitting the biphasic Hill equation to the experimental data obtained in the presence of various [InsP3] with Hact (activation Hill coefficient) fixed.
Sf9-InsP3R channel inactivation following InsP3-mediated activation
In all ligand conditions, channel activity invariably terminated in the continuous presence of InsP3 and absence of local change in [Ca2+]i (Figs 1 and 3A). In optimal ligand conditions, the mean channel activity duration (Ta) was ∼120 s. Similar apparent inactivation has been observed for nuclear membrane-patched InsP3R in Xenopus oocytes (Mak & Foskett, 1994, 1997) and COS-7 cells (Boehning et al. 2001a) but with significantly shorter Ta. In 10 μm InsP3, Ta of the Sf9-InsP3R was reduced in [Ca2+]i >1 μm, with reduction by over 10-fold at 89 μm Ca2+ (P < 0.05) (Fig. 3A). In subsaturating (33 nm) InsP3, Ta began to decrease in [Ca2+]i∼300 nm, substantially lower than that observed in saturating InsP3 (P < 0.05) (Fig. 3A). The [InsP3] and [Ca2+]i dependencies of Ta suggest that the observed inevitable termination of channel activity represents the InsP3R channels entering a true inactivated state instead of an experimental artefact.
Figure 3. Dependencies on [Ca2+]i and [InsP3] of InsP3R channel activity duration and recruitment.
A, channel activity duration. Points represent average of channel activity durations for multiple experiments (n > 10) in ligand conditions as shown. Asterisks (*P < 0.05, **P < 0.01) indicate data points that are statistically significantly different from the reference point (circled points). Red is for comparisons of data obtained in 10 μm InsP3. Blue is for comparison of data obtained in 33 nm InsP3. Green is for comparisons of data obtained in optimal 1 μm Ca2+i. Smooth curves were drawn by hand for clarity. B–D, ligand-dependent recruitment of InsP3R. B, channel detection probability (Pd) in ligand concentrations as shown. Total numbers of membrane patches obtained in each set of ligand conditions are displayed next to data points. C, mean number of active channels in membrane patches (NA) in ligand concentrations as shown. Asterisks (*P < 0.05, **P < 0.01) indicate data points that are statistically significantly different from the reference points (circled points). Red is for comparisons of data obtained in 10 μm InsP3. Blue is for comparisons of data obtained in 33 nm InsP3. Green is for comparisons of data obtained in optimal 750 nm Ca2+i. D, the product NAPo determined using data shown in Figs 2B and 3C, in ligand concentrations as shown. The smooth curves were hand drawn for clarity.
Nevertheless, without evidence of reversibility, the ligand dependence of the channel activity duration by itself is still insufficient to determine definitively whether the activity termination represents true channel inactivation. The orientation of the channel, with its ligand binding sites facing into the pipette solution, does not allow for wash out and re-addition of InsP3 to test whether the observed termination process is reversible. We therefore used an alternative protocol in which the channels in isolated nuclei were first exposed to InsP3 that had been added to the bath for variable amounts of time before the nuclei were patched. If channel activity termination observed in patched membranes is an intrinsic feature of the InsP3R, then pre-exposure of nuclei to InsP3 should reduce the number of active channels detected (NA) during patch clamping. Furthermore, if the termination process is reversible, the number of channels detected should increase again after the InsP3 is washed out of the bath. With the bath containing 300 nm free Ca2+ and 10 μm InsP3, and the patch pipette solution containing 10 μm InsP3 and 1 μm Ca2+, NA decreased with time until no channels were detected (NA= 0) by 900 s exposure to InsP3 (Fig. 4, protocol A). Channels could not be detected during the remainder of the 30 min exposure to bath InsP3, but NA recovered quickly upon washout of the bath InsP3. As control, NA in patches obtained from nuclei not pre-exposed to InsP3 remained constant over this period. This reversible disappearance of detectable channels indicates that the InsP3R undergoes true InsP3-dependent channel inactivation. We also examined the effects of pre-exposure to InsP3 in the presence of high Ca2+, because we expect from the channel duration measurements that the inactivation rate should be considerably faster. Here, nuclei were first exposed to 10 μm InsP3 in the presence of 300 nm Ca2+ for 2 min, to enable channel activation, and then the bath Ca2+ was raised to 49 μm in the continuous presence of InsP3, to accelerate inactivation rate (Fig. 4, protocol B). NA was rapidly reduced to zero within 2 min (approximately the temporal resolution of this approach), and then recovered when the bath was changed to one that lacked InsP3 and contained 300 nm Ca2+. This reversible disappearance of detectable channels again indicates that the InsP3R undergoes true InsP3-dependent channel inactivation, and the much more rapid kinetics in high Ca2+ agree with the shorter activity duration measurements observed in high [Ca2+] (Fig. 3A). The kinetics of inactivation measured by the InsP3-pre-exposure protocol were slower than those determined from the measurements of channel activity durations of patched nuclei in both 300 nm and 49 μm Ca2+, but the results of the two protocols are qualitatively similar, i.e. inactivation is several fold faster at 49 μmversus 300 nm Ca2+ in both protocols.
Figure 4. Reversible loss of detectable InsP3R channels by pre-exposure to InsP3.
Isolated nuclei were exposed to 10 μm InsP3 during a 30–34 min period in normal bath (300 nm Ca2+; protocol A) or in the presence of 49 μm Ca2+ (protocol B), as indicated. Nuclear patch clamping was performed at various times during and after the exposure of the nuclei to bath InsP3, with the pipette solution always containing 10 μm InsP3 and 1 μm Ca2+. The number of active channels observed in each patch (NA) is shown in the graph. The data are combined from three separate batches of isolated nuclei. Filled circles, NA in patches from nuclei exposed to bath InsP3 up to the time of patching; open circles, NA in patches from nuclei that had been previously exposed to InsP3, and subsequently washed free of bath InsP3 by multiple bath replacements before patching; diamonds, NA in patches from control nuclei exposed to bath solution containing 300 nm Ca2+ and no InsP3. Pre-exposure to bath InsP3 caused NA to decrease to zero, slowly in 300 nm bath Ca2+ and rapidly in 49 μm bath Ca2+. NA recovered to control levels in both protocols upon removal of InsP3 and restoration of normal bath conditions.
Probability of detecting Sf9-InsP3R channels in membrane patches
The conductance, activation and inactivation gating properties and their regulation by InsP3 and [Ca2+]i, and spatial distribution of the observed Sf9 InsP3R channels are all highly reminiscent of other InsP3R channels observed in nuclear patch clamp studies. In those previous studies, the probability of detecting an active InsP3R channel in the patched membrane (Pd) was generally low, i.e. most patches contained no active channels; and there was significant variability in Pd (Mak & Foskett, 1997; Mak et al. 2000; and unpublished observations in COS-7 cells). In contrast, active Sf9-InsP3R channels were detected in optimal ligand conditions in up to ∼80% of patches obtained from nuclei from different batches of cells. This consistency allowed us to examine the possible ligand dependence of Pd (Fig. 3B). In 10 μm InsP3 in low [Ca2+]i (50 nm), Pd was 0.47. That is, nearly half of all patches contained at least one active InsP3R channel. Pd increased to 0.81 when [Ca2+]i was raised to 500 nm. Between 0.5 and 7.5 μm, Pd remained high (∼0.6–0.8). When [Ca2+]i was increased beyond 10 μm, Pd gradually decreased. Thus, Pd varied as a biphasic function of [Ca2+]i. With 33 nm InsP3, Pd increased from 0.5 at [Ca2+]i= 100 nm to 0.67 at [Ca2+]i= 500 nm, and then decreased to 0.21 at [Ca2+]i= 10 μm (Fig. 3B). Pd was sensitive to [InsP3] as well as [Ca2+]i, with decreasing [InsP3] within the subsaturating range enhancing high [Ca2+]i reduction of Pd, reminiscent of the effect of suboptimal [InsP3] on high [Ca2+]i inhibition of channel gating (Fig. 2B).
To calculate Pd, a membrane patch containing multiple channels was given the same weight as one containing a single channel. However, it was noted that patches obtained under optimal ligand conditions usually contained several active channels, whereas those obtained under suboptimal conditions often contained only one. The relatively long channel lifetimes allowed the number of channels to be determined accurately under most conditions (see Methods). To more accurately characterize the number of channels activated, we defined NA as the total number of channels observed in all patches divided by the total number of patches obtained under that set of experimental conditions. In [Ca2+]i= 50 nm, in the presence of 10 μm InsP3, NA was 1.3 ± 0.3. That is, on average each patch contained 1.3 active InsP3R channels. In higher [Ca2+]i, more channels were detected in each patch, with maximum NA of 3.0 ± 0.4 at [Ca2+]i= 500 nm (P < 0.01). In 1 μm < [Ca2+]i < 8 μm, NA ranged between 2.2 and 2.8 (Fig. 3C). Further increases in [Ca2+]i were associated with reduced NA(P < 0.01). Compared with NA observed in 10 μm InsP3, consistently lower NA was observed in 33 nm InsP3 over all [Ca2+]i (P < 0.01). In [Ca2+]i= 110 nm, NA was 0.7 ± 0.2. It increased to 1.6 ± 0.3 at [Ca2+]i= 500 nm. Higher [Ca2+]i then decreased NA(P < 0.01), such that at [Ca2+]i= 7.5 μm, only 0.26 ± 0.10 channels per patch were detected.
The observation that NA was a function of stimulus strength was unexpected since it was anticipated that the entire channel population in a membrane patch would always become activated, albeit to different levels of activity (Po) depending on the strength of ligand activation. These differences in NA cannot be explained by under-estimation of the number of channels when Po was low, since Sf9-InsP3R channels were observed long enough for NA to be accurately determined (see Methods). Furthermore, since the observation period following exposure of the membrane patch to ligands was many tens of seconds, the inability of suboptimal ligand concentrations to engage all channels cannot be accounted for by stochastic probabilities of ligand binding to the channels. These results indicate that suboptimal ligand concentrations are insufficient to activate all InsP3R channels in a membrane patch that can be activated by optimal ligand concentrations. Therefore, ligand activation of InsP3R is associated with recruitment of channels to an activated state, as well as with increasing the Po of those channels that are recruited.
The total ion flux (J) associated with activation of the entire population of InsP3R in the ER membrane is defined by:
| 2 |
where γ is the single-channel conductance, N is the number of activated channels, and Po is the average single-channel open probability. The patch-clamp experiments were able to measure all three parameters independently, enabling the effects of ligand stimulation on the magnitude of InsP3R-mediated Ca2+ release from the ER store to be estimated. With γ being ligand independent, NAPo is a measure of the ligand-induced InsP3R-mediated Ca2+ release. As both parameters were sensitive to ligand concentrations, NAPo is a more accurate assessment of the consequences of ligand activation of the channel than either parameter alone. In saturating [InsP3], the dependence of NAPo on [Ca2+]i was biphasic: NAPo increased by over 10-fold as [Ca2+]i was increased from 50 to 500 nm (Fig. 3D), and then gradually decreased as [Ca2+]i was further increased. NAPo was also strongly dependent on [InsP3]. With [InsP3] reduced to 33 nm, the dependence of NAPo on [Ca2+]i was again biphasic, with maximal NAPo observed in [Ca2+]i≈ 0.5–1 μm, but peak NAPo was an order of magnitude lower than that observed in saturating [InsP3] (Fig. 3D).
Discussion
We characterized the single-channel permeation and gating properties of the endogenous Sf9 cell InsP3R in native ER membranes, and found them to be remarkably similar to those of its vertebrate homologues. Among various cell types used for nuclear patch-clamp studies of InsP3R (COS-7, Xenopus oocyte, DT40, and CHO; our published and unpublished results), functional InsP3R channels are most frequently and consistently detected using Sf9 cell nuclei, which is distinctly advantageous compared with the other systems in which fewer channels are detected less reliably. We exploited this to demonstrate, for the first time, two important features of InsP3R channel regulation. First, ligand activation of InsP3R channel gating inevitably initiates an inactivation mechanism that terminates gating. Second, the number of InsP3R channels that are recruited to activation is a graded function of ligand concentration.
Functional properties of the endogenous Sf9 InsP3R
We utilized a nuclear isolation protocol that avoids purification and reconstitution procedures, enabling the InsP3R channels to remain in their native membrane environment, preserving specific lipid interactions during our experiments. Previously, microsomes obtained from transduced Sf9 cells were used as a source of recombinant InsP3R in biochemical studies (Cardy et al. 1997) and lipid bilayer reconstitution studies (Tu et al. 2002, 2004; Srikanth et al. 2004). Whereas it was reported in those studies that endogenous Sf9 InsP3R were not detected biochemically or functionally, the InsP3 signalling system is present and functional in Sf9 cells (Ross et al. 1994; Knight et al. 2003; Knight & Grigliatti, 2004), and our studies here have revealed that the Sf9-InsP3R is functionally expressed at a high levels in the outer nuclear membrane of freshly isolated nuclei. The apparent discrepancy may reflect a substantially higher sensitivity of detection of InsP3R function in patch clamp electrophysiology, enrichment of InsP3R in the nuclear envelope, or use of antibodies in biochemical studies that only poorly recognized the Sf9 InsP3R.
Under our experimental conditions, with symmetrical K+ as the charge carrier (Mak et al. 1998), the Sf9-InsP3R channel exhibited a linear current–voltage relation, similar to that observed for nuclear-patched endogenous Xenopus InsP3R-1, rat cerebellar InsP3-R (Marchenko et al. 2005) and recombinant rat types 1 and 3 InsP3R (Mak & Foskett, 1998; Boehning et al. 2001a; Mak et al. 2001); however, the single-channel conductance of ∼480 pS of the Sf9 InsP3R channel is higher than the ∼360 pS conductance of the vertebrate channels. This may be due to the membrane environment of the Sf9 cell. Alternatively, it may be due to differences at the molecular level between the insect and mammalian InsP3R pore sequences. Although the amino acid sequence of the Sf9-InsP3R is not known, all known invertebrate InsP3R pore selectivity filter sequences (BLAST search: Drosophila, Caenorhabditis elegans, Panulirus, Anopheles, Asterina, Apis, Strongylocentrotus, Aplysia and Tetrahymena) contain a GGIGD motif, whereas the vertebrate sequence is GGVGD. By site-directed mutagenesis of rat InsP3R-1, it was previously demonstrated that this sequence is involved in ion conductance and selectivity (Boehning et al. 2001b). A mutant mammalian InsP3R channel with the pore Val replaced with Ile to resemble the invertebrate InsP3R, had a higher conductance (490 ± 13 pS) (Boehning et al. 2001b), close to that (∼480 pS) of the Sf9-InsP3R measured here. Thus, other invertebrate InsP3R channels may also have higher single-channel conductance than their mammalian counterparts. In contrast, the ion permeability selectivity of the Sf9 channel is very similar to that of its mammalian homologs, with the channel cation selective (PK:PCl∼5) with ∼10-fold higher selectivity for Ca2+ than K+ (Mak & Foskett, 1994, 1998; Mak et al. 2000; Boehning et al. 2001a; Marchenko et al. 2005).
We examined the gating properties of the Sf9-InsP3R under a wide range of ligand concentrations using experimental conditions similar to those used to study other InsP3R channels in native ER membranes (Mak et al. 1998; Boehning et al. 2001a). The most fundamental conclusion is that the ligand regulation properties of the insect Sf9-InsP3R have remarkable similarities to those of the Xenopus type 1 and rat types 1 and 3 channels studied previously (Mak et al. 1998, 2001; Boehning et al. 2001a). Like the vertebrate channels, the Sf9-InsP3R has a high maximum Po of 0.7–0.8, and the Sf9 InsP3R channel Po has a biphasic dependence on [Ca2+]i, with the ranges of [Ca2+]i over which the channel was activated or inhibited similar to those of the vertebrate channels recorded under similar conditions of [InsP3] and [ATP]. Also like the vertebrate channels, the Sf9-InsP3R channel Po remained at high levels over a broad range of [Ca2+]i, with inhibition by high [Ca2+]i not pronounced until [Ca2+]i > 10 μm in the presence of saturating [InsP3]. However, inhibition of the Sf9-InsP3R channel by high [Ca2+]i is less cooperative than the vertebrate InsP3R isoforms (Mak et al. 2001). Consequently, Po is already <0.1 in the vertebrate channels at [Ca2+]i= 90 μm, whereas Sf9-InsP3R Po is still about 0.2 at [Ca2+]i= 90 μm (Fig. 2B).
The main effect of reducing [InsP3] to subsaturating levels (∼100 nm) on the Sf9-InsP3R channel activity was to increase the sensitivity of the channel Po to inhibition by high [Ca2+]i. In contrast, Ca2+ activation was not significantly affected by [InsP3]. This is highly reminiscent of InsP3 regulation of endogenous Xenopus InsP3R-1 (Mak et al. 1998) and recombinant rat InsP3R-3 (Mak et al. 2001). Nevertheless, the Sf9-InsP3R has a lower effective sensitivity to InsP3 and the relief of Ca2+ inhibition by InsP3 has less apparent cooperativity in Sf9-InsP3R compared with the vertebrate channels (Fig. 2B). Both Sf9-InsP3R and Xenopus InsP3R-1 have low channel activity in 10 nm InsP3, but whereas Xenopus InsP3R-1 is already fully activated by 100 nm InsP3, Sf9-InsP3R responds to InsP3 over a significantly broader range of concentrations: increasing [InsP3] from 100 nm to 1 μm gave rise to further relief of Ca2+ inhibition of Po in Sf9-InsP3R (Fig. 2B).
Finally, the kinetic basis for Sf9-InsP3R channel activation by InsP3 and Ca2+ involves primarily modulation of tc. The to value remained within a relatively narrow range, whereas tc varied over an order of magnitude over a range of [Ca2+]i from 50 nm to 90 μm. Thus, the [Ca2+]i dependence of channel tc is mostly responsible for the channel Poversus[Ca2+]i relation. This is similar to the effect of [Ca2+]i on the Xenopus InsP3R-1 and rat InsP3R-3 channels (Mak et al. 1998, 2001).
The broad similarities of ligand regulation among the Sf9-InsP3R channel and two types of vertebrate InsP3R channels suggest that common conserved mechanisms underlie ligand regulation of all InsP3Rs. A previous study of the Drosophila InsP3R also concluded that its functional properties were similar to those of mammalian InsP3Rs (Swatton et al. 2001). The [Ca2+]i and [InsP3] dependencies of the Sf9-InsP3R channel activity can be adequately described by an allosteric MWC-based model (Fig. 2C) that accounted for the ligand regulation of the Xenopus InsP3R-1 and rat InsP3R-3 channels observed in previous nuclear patch-clamp studies (Mak et al. 2003a). The remarkable functional similarity of Sf9-InsP3R with the endogenous Xenopus and recombinant rat InsP3Rs may be a consequence of the high sequence homology among different InsP3R isoforms from various species.
InsP3-induced InsP3R channel inactivation
In single-channel studies, gating activities of Sf9-InsP3R (the present study) and vertebrate InsP3R (Mak & Foskett, 1997; Mak et al. 2000; Boehning et al. 2001a) recorded in native ER membranes inevitably abruptly terminate after InsP3 activation. Whereas the previous vertebrate single-channel studies suggested that the abrupt termination was likely to be due to channel inactivation, it was impossible to rule out non-physiological artefacts associated with patching, for example collisions of the channels with the walls of the glass pipette. However, the present results, by demonstrating both InsP3 as well as Ca2+ dependencies of the channel activity duration, and the reversibility of the activity termination, clearly establish this process as true channel inactivation. Previously, the presence of InsP3-induced InsP3R inactivation has been controversial. Pre-exposure of permeabilized hepatocytes to InsP3 under conditions of constant [Ca2+]i was followed by a time-dependent reduction of InsP3R-mediated Ca2+ release (Hajnóczky & Thomas, 1994, 1997; Dufour et al. 1997; Marchant & Taylor, 1998) as well as transformation of the receptor from a low-affinity active state to a high-affinity, desensitized state (Coquil et al. 1996; Marchant & Taylor, 1998), consistent with InsP3-mediated InsP3R inactivation. Biphasic kinetics of Ca2+ release rates in intact cells have been observed in many studies (reviewed in Bootman, 1994; Missiaen et al. 1994; Parys et al. 1996; Taylor, 1998), but either inactivation has been ruled out as a mechanism for release termination (Taylor & Potter, 1990; Oldershaw et al. 1992; Hirose & Iino, 1994; Parys et al. 1995; Combettes et al. 1996; Beecroft & Taylor, 1997) or other types of mechanisms have been invoked (reviewed in Bootman, 1994; Missiaen et al. 1994; Parys et al. 1996; Taylor, 1998). In rapid perfusion protocols using isolated microsomes or permeabilized cells that attempted to maintain [InsP3] and [Ca2+]i constant, transient Ca2+ release kinetics indicated that channel inactivation or partial inactivation played a role in the decay of release rates (Champeil et al. 1989; Finch et al. 1991a; Combettes et al. 1994; Wilcox et al. 1996; Dufour et al. 1997; Marchant & Taylor, 1998; Adkins et al. 2000), but questions have been raised regarding the efficacy of maintaining constant conditions in these protocols (Taylor, 1998). Observations of refractory periods following either global (Khodakhah & Ogden, 1995; Carter & Ogden, 1997; Ogden & Capiod, 1997) or more focal (Parker & Ivorra, 1990a; McCarron et al. 2004) InsP3-mediated Ca2+ release in intact cells are also consistent with channel inactivation in intact cells. However, in rapid perfusion and cell studies, it has remained unclear whether release termination or decay is caused by a true intrinsic InsP3-induced inactivation process, possibly accelerated by high [Ca2+]i, or whether it is due to Ca2+-feedback inhibition of channel gating. Furthermore, it has not been established if apparent inactivation as a consequence of pre-exposure to Ca2+ before InsP3 exposure (Parker & Ivorra, 1990a; Payne et al. 1990; Finch et al. 1991b; Combettes et al. 1994; Oancea & Meyer, 1996; Ogden & Capiod, 1997; Swatton & Taylor, 2002; McCarron et al. 2004) is related to apparent inactivation as a result of InsP3 binding.
The single-channel studies reported here, performed under conditions of constant ligand concentrations and ER luminal conditions, indicate that InsP3-induced inactivation is an intrinsic property of the single InsP3R channel. Although the kinetics of apparent inactivation observed in superfusion experiments (Finch et al. 1991b; Combettes et al. 1994; Dufour et al. 1997) and in intact cells in response to photo-release of InsP3 (Khodakhah & Ogden, 1995) were rapid (<1 s), the inactivation kinetics reported here are slower. However, the kinetics of inactivation for InsP3R in nuclear patch-clamp studies observed here (Ta∼ 10–100 s) and previously (Ta∼ 20–30 s; Mak & Foskett, 1997), and here in the InsP3-pre-exposure protocols, are similar to those estimated by ER permeability measurements in permeabilized hepatocytes (Hajnóczky & Thomas, 1994), as well as to the kinetics of InsP3-induced increases in InsP3 affinity of an apparently desensitized InsP3R in cerebellar microsomes (Coquil et al. 1996) (t½∼15–45 s) and of the transient fast phase of Ca2+ release in response to initial exposure to InsP3 in permeabilized and intact cells (e.g. see Muallem et al. 1989; Meyer & Stryer, 1990; Taylor & Potter, 1990). Of note, InsP3R inactivation observed in permeabilized hepatocytes (Hajnóczky & Thomas, 1994), which had kinetics similar to those observed in our single channel studies, was shown to account for release termination associated with [Ca2+]i oscillations (Hajnóczky & Thomas, 1997). Together these results suggest that the kinetics of inactivation observed in single-channel studies are of physiological relevance for [Ca2+]i signalling observed in cells. However, it remains unclear whether distinct inactivation kinetics observed in different studies reflects methodological differences, distinct types of inactivation or inhibition, or a physiologically relevant range of kinetics of a common inactivation mechanism.
Our results here demonstrate that the kinetics of single channel inactivation are sensitive to both [InsP3] and [Ca2+]i. Inactivation was observed for every activated channel, even under conditions of saturating [InsP3] and very low [Ca2+]i (50 nm). Thus, inactivation appears to be fatefully linked to activation, but the rate of channel inactivation was accelerated at lower [InsP3] and higher [Ca2+]i, with channel activity duration reduced by over an order of magnitude between 1 and 89 μm Ca2+, with average channel duration in 89 μm Ca2+ reduced to under 10 s. In addition to ligand concentrations, other factors may also be important in determining channel inactivation kinetics. For example, the rate of channel inactivation of the Xenopus InsP3R (Mak & Foskett, 1997) was faster than that the Sf9-InsP3R studied here under identical ligand conditions. Furthermore, the rate of the Sf9-InsP3R inactivation was faster in the patched membranes than in the InsP3-pre-exposure protocols. Thus, factors such as membrane tension, membrane lipid composition and associated proteins may also contribute to the rate of InsP3R channel inactivation. It is interesting that inactivation has not been reported for reconstituted InsP3R channels recorded in artificial planar bilayer membranes. Our results suggest that inactivation rates of single InsP3R channels are regulated by ligand concentrations and possibly by additional, including cell-type-specific, factors as well, consistent with a common regulated mechanism of inactivation. Importantly, however, additional studies will be required to determine whether inactivation observed in single channel studies contributes to Ca2+ release termination in cells.
Inactivation versus inhibition
Because the curves that describe the transient kinetics of Ca2+ release observed in cells and rapid perfusion experiments are reminiscent of the biphasic curves that describe the [Ca2+]i dependencies of steady-state channel Po, there has been a tendency in the literature to equate the two and to account for release termination by effects of high [Ca2+]i on single-channel gating activity. Consequently, the terms ‘inhibition’ and ‘inactivation’ and ‘partial inactivation’ have sometimes been used to refer to similar phenomena. However, as pointed out (Sneyd et al. 2004), the bell-shaped or otherwise biphasic shape of the steady state Poversus[Ca2+]i curve has ‘very little, if anything’, to do with the fact that the InsP3R exhibits complex rapid kinetic behaviours. Our single-channel studies of Sf9-InsP3R demonstrate that channel inactivation occurs over all [Ca2+]i studied, with the rate of inactivation not necessarily correlated with channel Po. Of note, it was previously observed that Xenopus oocyte InsP3R channels that lack high [Ca2+]i inhibition of channel Po (as a result of an experimental manoeuvre) nevertheless still inactivated (Mak et al. 2003b). Thus, InsP3R inhibition by high [Ca2+]i, and InsP3-induced inactivation of the InsP3R, are distinct processes. The [Ca2+]i dependencies of channel Po inhibition and inactivation are likely to be mediated by distinct Ca2+-binding sites, as discussed in more detail below. Consequently, we propose that the term ‘inhibition’ be used to refer to the effects of high [Ca2+]i on steady-state channel Po, whereas ‘inactivation’ be used to refer to the time-dependent loss of the capacity of the channel to open even in the presence of constant [InsP3] and [Ca2+]i.
Ligand stimulation by both InsP3R channel recruitment and activation
We found that the average number of channels activated in a membrane patch (NA) depends on the strength of ligand activation. NA had a biphasic dependence on [Ca2+]i, and was smaller when InsP3 was reduced to suboptimal concentrations over the entire range of [Ca2+] examined (Fig. 3C). These differences in NA cannot be explained by underestimation of the number of channels when Po was low, since the channel current records were long enough for NA to be accurately determined. Notably, the ligand concentration dependencies of Po and NA are similar. Thus, the product NAPo is a sensitive function of both [Ca2+]i and [InsP3] (Fig. 3D), and NAPo quantitatively accounts for the magnitude of the Ca2+ flux associated with InsP3-mediated [Ca2+]i signals in cells (eqn (2)) because channel conductance is constant. Our single-channel results therefore suggest that the magnitudes of InsP3R-mediated Ca2+ signals in cells are determined by both channel recruitment as well as single-channel activity.
The observed ligand dependence of the number of channels that become recruited into the activated state is unexpected. Because the channels were observed for tens of seconds during their exposure to the ligands, graded channel recruitment cannot be accounted for by stochastic probabilities of ligand binding to the channels. Thus, it was anticipated that the entire channel population in a membrane patch would always become engaged, albeit to different levels of activity (Po) depending on the strength of ligand stimulation. The graded recruitment of the number of activated channels seemingly implies that the InsP3R channels in Sf9 cell nuclei have heterogeneous sensitivities to both [InsP3] and [Ca2+]i. Nevertheless, it seems unlikely that heterogeneous ligand sensitivity exists among the channels. InsP3R isoform diversity cannot be involved since no invertebrate has been demonstrated to express more than one InsP3R isoform. Furthermore, it is unlikely that heterogeneous covalent modifications could account for the observed graded responsiveness to such a wide range of both ligands in nuclei from different batches of cells. In addition, possible local determinants of channel activity such as luminal [Ca2+] or channel density can be ruled out in the system studied here.
A qualitative model that accounts for ligand dependencies of channel recruitment and inactivation
A simple qualitative kinetic scheme (Fig. 5) involving channel sequestration can account for the observed ligand-dependent channel recruitment without invoking intrinsic channel heterogeneity. The model assumes that InsP3R channels can become sequestered in a non-active S state by the binding of Ca2+ to a sequestration site. In the absence of InsP3, the regular closed C state of the channel is in equilibrium with the S state, with the C state being favoured (Fig. 5A). When the channel is exposed to InsP3, InsP3 binding to the channel enables the channel to enter activated A states (including the A* state in which the channel is open to conduct ion flux, and the closed A′ state). A competition ensues in which the probability of the channel becoming activated, which depends on [Ca2+]i according to the previously developed MWC-based allosteric model (Mak et al. 2003a), is countered by the probability of the channel becoming sequestered into the non-active S state (Fig. 5B). Ca2+ concentrations that favour activation of the InsP3R channel (optimal [Ca2+]i) decrease the chance of the channel being sequestered before activation, increasing NA; in contrast, suboptimal [Ca2+]i (subactivating [Ca2+]i or high inhibitory [Ca2+]i) that are less favourable for channel activation increase the chance that the channel is sequestered into the S state and therefore reduce NA. Consequently, the [Ca2+]i dependence of NA is biphasic and similar to the [Ca2+]i dependence of channel Po under all [InsP3]. Once an InsP3-bound channel enters the S state, it is stabilized there because the affinities of its InsP3-binding sites and the sequestration Ca2+-binding sites are so high that it cannot escape from that state in the continuous presence of InsP3. For any particular [Ca2+]i, NA is lower in subsaturating [InsP3] than saturating [InsP3] because lower [InsP3] allows more channels to be siphoned off into the S state by Ca2+ binding to the sequestration site before InsP3 can bind to the channel. Subsequent binding of InsP3 to the channel stabilizes it in the S state.
Figure 5. Schematic diagram of a model to account for ligand dependencies of graded InsP3R channel recruitment and InsP3-induced inactivation.
A, C, S and I represent activated, closed, sequestered and inactivated InsP3R channel states, respectively. Subscripts indicate state of InsP3-binding. The InsP3R channel states are connected by arrows, with triangular arrowheads representing Ca2+ and InsP3 binding to the channel as indicated. ○, Ca2+ binding to sequestration site; §, Ca2+ binding to inactivation site. A bigger arrowhead indicates the side of the reaction or transition that the equilibrium favours. Reactions or transitions with thicker arrow shafts have higher rates. The activated InsP3-bound A state can gate between channel open (A*) and closed (A′) conformations through ligand-independent transitions represented by arrows with chevron arrowheads. Equilibrium between InsP3R channels in C and A states is represented by the open arrow marked with MWC, discussed in detail in (Mak et al. 2003a). See text for a full description of the model.
According to this scheme, termination of channel activity due to inactivation is caused by the channel going from the closed activated state (A′ state) to an inactive I state as the result of Ca2+ binding to an inactivation site (Fig. 5C). Because the channel only enters the I state from the A′ state, conditions that reduce channel Po (high inhibitory [Ca2+]i and subsaturating [InsP3]) increase the chance that the channel is in A′ state, and therefore accelerate channel inactivation. In subactivating [Ca2+]i, channel Po is low but channel inactivation is not accelerated because this effect is countered by the reduction of the rate of Ca2+ binding to the inactivation site due to lower [Ca2+]i. Like the S state, the channel in the I state has high affinities for InsP3 and for Ca2+ in the inactivation site, which prevents the channel from escaping from the I state for as long as InsP3 is present. Recovery from inactivation is only possible by InsP3 dissociation as InsP3 is removed. This accounts for the observed reversibility from inactivation upon wash-out of InsP3 (Figs 4 and 5).
There are obvious similarities in the properties of channel sequestration, which results in a certain fraction of available InsP3R channels not being activated despite the continuous presence of InsP3, and channel inactivation, which results in InsP3-induced termination of channel activity. In both processes, the InsP3R channel ends up in a state from which it cannot escape in the continuous presence of InsP3. Both processes are regulated by [InsP3] and [Ca2+]i. And, in our scheme, both processes involve Ca2+ binding to the InsP3R channel. However, the time scales of the two processes are very different. Channel sequestration must occur on a time scale shorter than the activation of the InsP3R channel after it is exposed to InsP3, within ∼1–5 s (from the time the patch pipette makes contact with the membrane and the channels in the isolated membrane patch are exposed to InsP3, to the time when the gigaohm seal forms to allow recording of channel currents). In contrast, channel inactivation occurs at least an order of magnitude slower, ∼20–100 s. Therefore, it is likely that sequestration and inactivation are distinct processes, involving two distinct Ca2+-binding sites.
This simple qualitative modelling of our single-channel results indicates that InsP3-induced channel inactivation (I states in Fig. 5) from active conformations may play a role in terminating Ca2+ release. Here, we now propose that channel sequestration before it releases Ca2+, plays a role in tuning the number of release channels involved in responses in cells. Interestingly, this model is reminiscent of a kinetic model developed by Sneyd & Dufour (2002), that closely followed the ideas of Taylor (Marchant & Taylor, 1997; Adkins & Taylor, 1999), to account for Ca2+ release kinetics from hepatic microsomes, and of another kinetic model developed by Dawson et al. (2003) that was adapted from a kinetic scheme of ryanodine receptor adaptation (Sachs et al. 1995). In both kinetic models, inactivated InsP3R states were also accessible from either open or closed channel states. Although our simple model cannot predict transient kinetic features, and the two kinetic models either did not or could not account well for steady-state single-channel data, it is interesting that the model here, based primarily on equilibria among channel conformations that accounts for steady-state single-channel activities, has a similar transition diagram with one based on rate constants of ligand binding reactions that accounts for dynamic Ca2+ release by a population of InsP3R channels. Future studies of the rapid kinetic responses of single InsP3R channels to step changes in ligand concentrations may provide data necessary to extend the insights derived from models of steady-state channel gating to understanding kinetic features that underlie dynamic [Ca2+]i signals in cells.
Physiological implications of InsP3R channel recruitment
An intriguing aspect of InsP3R-mediated Ca2+ release is that it can be graded in response to incremental levels of extracellular agonist or [InsP3]. Many mechanisms that have been proposed to account for graded release invoke heterogeneity of either the Ca2+ stores or the release channels. Thus, different Ca2+ stores with distinct sensitivities to InsP3 (Muallem et al. 1989; Kindman & Meyer, 1993) or different densities of InsP3R channels (Meyer & Stryer, 1990; Hirose & Iino, 1994) have been proposed to account for graded Ca2+ release. Alternately, it was hypothesized that graded responses are due to progressive recruitment of independent, so-called elementary release units within a continuous Ca2+ store, including Ca2+ puffs and sparks (Berridge, 1997). Although it has been suggested that InsP3 or Ca2+ might regulate the stochastic probability of triggering an InsP3R-mediated Ca2+ release event (Yao et al. 1995; Horne & Meyer, 1997), localized Ca2+ release events are observed from ‘eager sites’ even under conditions of uniform cytoplasmic [InsP3] and [Ca2+] (Yao et al. 1995), suggesting apparent heterogeneity of InsP3 sensitivity among release sites. As pointed out in (Bootman & Berridge, 1995), a key issue to account for graded Ca2+ release is to understand the mechanisms involved in ‘progressive recruitment’ of local release events.
The results of the present study now suggest one possible mechanism that does not require channel or store heterogeneity. We observed that single InsP3R channels are recruited into active conformations as a graded function of stimulus intensity, in a uniform population of InsP3R channels in membrane patches with apparent constant channel density and luminal conditions, ruling out factors invoked in many previously proposed mechanisms. Indeed, our results strongly suggest that graded recruitment of channels is an intrinsic property of the InsP3R channel itself, consistent with graded Ca2+ release having been observed in vesicles containing purified InsP3R protein (Ferris et al. 1992; Hirota et al. 1995) and in a wide variety of cell types (Bootman, 1994; Missiaen et al. 1994; Parys et al. 1996; Taylor, 1998).
It was previously proposed that high-[Ca2+]i feedback inhibition of InsP3R gating, tuned by [InsP3], could provide a mechanism to grade release with stimulus intensity (Mak et al. 1998). Because high [Ca2+]i inhibits InsP3R channel gating by stabilizing the channel closed state (Fig. 2D) (Mak et al. 1998, 2001), Ca2+ release by one InsP3R channel can feed back to stabilize the closed channel state of other channels in the immediate proximity, preventing them from opening. Because InsP3 decreases the sensitivity of the channel to high [Ca2+]i inhibition (Mak et al. 1998, 2001), this could be a mechanism to grade the number of release channels activated, and consequently the amount of Ca2+ released. Thus, two distinct mechanisms exist that can grade the number of InsP3R channels that become activated in response to InsP3, one intrinsic to the channel itself mediated by sequestration processes revealed in this study, and another involving communication among channels by released Ca2+-mediated inhibition. Both mechanisms enable the initial rate of Ca2+ release to be tuned to the strength of ligand activation even in a homogenous population of channels. In addition, the activities of the recruited channels (Po) are also tuned to the strength of ligand activation. Simultaneous multiple channel recruitment and gating activation mechanisms enables Ca2+ release to be regulated by [InsP3] in a highly cooperative manner. It has been suggested that graded recruitment of release events involves the progressive recruitment of more signalling units that individually release Ca2+ in an all-or-none manner (Bootman & Berridge, 1995; Horne & Meyer, 1997). However, our single-channel results would suggest that both the recruitment of release sites (by channel recruitment) as well as the activity of each site (by Po effects on the activated channels and by recruitment of additional channels at each site through CICR) are graded functions of stimulus intensity, in agreement with imaging observations in oocytes (Callamaras & Parker, 1998; Sun et al. 1998) and HeLa cells (Thomas et al. 1998). Importantly, both features appear to be intrinsic to the InsP3R channel itself.
Graded recruitment of channels and inactivation of activated channels can provide a mechanism to account for quantal Ca2+ release observed in many studies. Our single-channel results are reminiscent of InsP3-induced incremental inactivation observed in permeabilized hepatocytes, in which pre-exposure to submaximal [InsP3] reduced the responses to subsequent higher [InsP3] in direct proportion to the amount of release induced by the pre-exposure to InsP3 (Hajnóczky & Thomas, 1994). Our results can account for this observation by recruitment of only a fraction of the total channel population by submaximal [InsP3] and subsequent inactivation of those recruited activated channels. However, our model cannot account for the InsP3 responsiveness of the remaining channels. Similarly, whereas our model can explain why sustained exposure to submaximal levels of agonists can only mobilize a fraction of total releasable Ca2+ in a cell, it cannot account for observations that subsequently increasing [InsP3] can release more Ca2+, thereby enabling Ca2+ stores to act as increment detectors of stimuli (Muallem et al. 1989; Meyer & Stryer, 1990; Bootman, 1994; Parys et al. 1996). Increment detection (Meyer & Stryer, 1990), unlike inactivation, enables the channels to retain full responsiveness to InsP3, whereas our proposed model of quantal Ca2+ release relies on inactivation from both open and closed channel states. Nevertheless, it is important to note that our model here is a simplified one, in particular because it largely ignores the tetrameric structure of the channel. It is possible that a more detailed scheme that accounted for ligand binding to each of the four monomers in a tetrameric channel could provide richer behaviours, for example to enable channels in inactivated states to retain the capacity to escape to open channel states in response to higher [InsP3]. Alternately, it is possible that factors not intrinsic to the channel, such as heterogeneous Ca2+ stores and InsP3R channels, luminal Ca2+ regulation, and InsP3R channel heterogeneity and density, play a role in enabling Ca2+ release to exhibit properties of increment detection. Thus, until studies have been performed that demonstrate that increment detection is a property of single InsP3R channels, development of more complicated models is unwarranted.
Supplementary Material
Acknowledgments
We thank John Pearson for helpful discussions of some iterations of various models. Supported by NIH grants (J.K.F.) and an American Heart Association Fellowship (C.W.).
References
- Adkins CE, Taylor CW. Lateral inhibition of inositol 1,4,5-trisphosphate receptors by cytosolic Ca2+ Curr Biol. 1999;9:1115–1118. doi: 10.1016/s0960-9822(99)80481-3. [DOI] [PubMed] [Google Scholar]
- Adkins CE, Wissing F, Potter BV, Taylor CW. Rapid activation and partial inactivation of inositol trisphosphate receptors by adenophostin A. Biochem J. 2000;352:929–933. [PMC free article] [PubMed] [Google Scholar]
- Beecroft MD, Taylor CW. Incremental Ca2+mobilization by inositol trisphosphate receptors is unlikely to be mediated by their desensitization or regulation by luminal or cytosolic Ca2+ Biochem J. 1997;326:215–220. doi: 10.1042/bj3260215. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Berridge MJ. Inositol trisphosphate and calcium signalling. Nature. 1993;361:315–325. doi: 10.1038/361315a0. [DOI] [PubMed] [Google Scholar]
- Berridge MJ. Elementary and global aspects of calcium signalling. J Physiol. 1997;499:291–306. doi: 10.1113/jphysiol.1997.sp021927. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Boehning D, Joseph SK, Mak D-OD, Foskett JK. Single-channel recordings of recombinant inositol trisphosphate receptors in mammalian nuclear envelope. Biophys J. 2001a;81:117–124. doi: 10.1016/s0006-3495(01)75685-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Boehning D, Mak D-OD, Foskett JK, Joseph SK. Molecular determinants of ion permeation and selectivity in inositol 1,4,5-trisphosphate receptor Ca2+ channels. J Biol Chem. 2001b;276:13509–13512. doi: 10.1074/jbc.C100094200. [DOI] [PubMed] [Google Scholar]
- Bootman MD. Quantal Ca2+ release from InsP3-sensitive intracellular Ca2+ stores. Mol Cell Endocrinol. 1994;98:157–166. doi: 10.1016/0303-7207(94)90134-1. [DOI] [PubMed] [Google Scholar]
- Bootman MD, Berridge MJ. The elemental principles of calcium signaling. Cell. 1995;83:675–678. doi: 10.1016/0092-8674(95)90179-5. [DOI] [PubMed] [Google Scholar]
- Bootman MD, Berridge MJ. Subcellular Ca2+ signals underlying waves and graded responses in HeLa cells. Curr Biol. 1996;6:855–865. doi: 10.1016/s0960-9822(02)00609-7. [DOI] [PubMed] [Google Scholar]
- Bootman MD, Lipp P, Berridge MJ. The organisation and functions of local Ca2+ signals. J Cell Sci. 2001;114:2213–2222. doi: 10.1242/jcs.114.12.2213. [DOI] [PubMed] [Google Scholar]
- Bootman M, Niggli E, Berridge M, Lipp P. Imaging the hierarchical Ca2+ signalling system in HeLa cells. J Physiol. 1997;499:307–314. doi: 10.1113/jphysiol.1997.sp021928. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Callamaras N, Parker I. Caged inositol 1,4,5-trisphosphate for studying release of Ca2+ from intracellular stores. Methods Enzymol. 1998;291:380–403. doi: 10.1016/s0076-6879(98)91024-2. [DOI] [PubMed] [Google Scholar]
- Cardy TJ, Traynor D, Taylor CW. Differential regulation of types-1 and -3 inositol trisphosphate receptors by cytosolic Ca2+ Biochem J. 1997;328:785–793. doi: 10.1042/bj3280785. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Carter TD, Ogden D. Kinetics of Ca2+ release by InsP3 in pig single aortic endothelial cells: evidence for an inhibitory role of cytosolic Ca2+ in regulating hormonally evoked Ca2+ spikes. J Physiol. 1997;504:17–33. doi: 10.1111/j.1469-7793.1997.00017.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Champeil P, Combettes L, Berthon B, Doucet E, Orlowski S, Claret M. Fast kinetics of calcium release induced by myo-inositol trisphosphate in permeabilized rat hepatocytes. J Biol Chem. 1989;264:17665–17673. [PubMed] [Google Scholar]
- Combettes L, Cheek TR, Taylor CW. Regulation of inositol trisphosphate receptors by luminal Ca2+ contributes to quantal Ca2+ mobilization. EMBO J. 1996;15:2086–2093. [PMC free article] [PubMed] [Google Scholar]
- Combettes L, Hannaert-Merah Z, Coquil J-F, Rousseau C, Claret M, Swillens S, Champeil P. Rapid filtration studies of the effect of cytosolic Ca2+ on inositol 1,4,5-trisphosphate-induced 45Ca2+ release from cerebellar microsomes. J Biol Chem. 1994;269:17561–17571. [PubMed] [Google Scholar]
- Coquil JF, Mauger JP, Claret M. Inositol 1,4,5-trisphosphate slowly converts its receptor to a state of higher affinity in sheep cerebellum membranes. J Biol Chem. 1996;271:3568–3574. doi: 10.1074/jbc.271.7.3568. [DOI] [PubMed] [Google Scholar]
- Dawson AP, Lea EJ, Irvine RF. Kinetic model of the inositol trisphosphate receptor that shows both steady-state and quantal patterns of Ca2+ release from intracellular stores. Biochem J. 2003;370:621–629. doi: 10.1042/BJ20021289. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dufour JF, Arias IM, Turner TJ. Inositol 1,4,5-trisphosphate and calcium regulate the calcium channel function of the hepatic inositol 1,4,5-trisphosphate receptor. J Biol Chem. 1997;272:2675–2681. doi: 10.1074/jbc.272.5.2675. [DOI] [PubMed] [Google Scholar]
- Ferris CD, Cameron AM, Huganir RL, Snyder SH. Quantal calcium release by purified reconstituted inositol 1,4,5-trisphosphate receptors. Nature. 1992;356:350–352. doi: 10.1038/356350a0. [DOI] [PubMed] [Google Scholar]
- Finch EA, Turner TJ, Goldin SM. Calcium as a coagonist of inositol 1,4,5-trisphosphate-induced calcium release. Science. 1991a;252:443–446. doi: 10.1126/science.2017683. [DOI] [PubMed] [Google Scholar]
- Finch EA, Turner TJ, Goldin SM. Subsecond kinetics of inositol 1,4,5-trisphosphate-induced calcium release reveal rapid potentiation and subsequent inactivation by calcium. Ann N Y Acad Sci. 1991b;635:400–403. doi: 10.1111/j.1749-6632.1991.tb36509.x. [DOI] [PubMed] [Google Scholar]
- Hajnóczky G, Thomas AP. The inositol trisphosphate calcium channel is inactivated by inositol trisphosphate. Nature. 1994;370:474–477. doi: 10.1038/370474a0. [DOI] [PubMed] [Google Scholar]
- Hajnóczky G, Thomas AP. Minimal requirements for calcium oscillations driven by the IP3 receptor. EMBO J. 1997;16:3533–3543. doi: 10.1093/emboj/16.12.3533. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hirose K, Iino M. Heterogeneity of channel density in inositol 1,4,5-trisphosphate-sensitive Ca2+ stores. Nature. 1994;372:791–794. doi: 10.1038/372791a0. [DOI] [PubMed] [Google Scholar]
- Hirota J, Michikawa T, Miyawaki A, Furuichi T, Okura I, Mikoshiba K. Kinetics of calcium release by immunoaffinity-purified inositol 1,4,5-trisphosphate receptor in reconstituted lipid vesicles. J Biol Chem. 1995;270:19046–19051. doi: 10.1074/jbc.270.32.19046. [DOI] [PubMed] [Google Scholar]
- Horne JH, Meyer T. Elementary calcium-release units induced by inositol trisphosphate. Science. 1997;276:1690–1693. doi: 10.1126/science.276.5319.1690. [DOI] [PubMed] [Google Scholar]
- Iino M. Biphasic Ca2+ dependence of inositol 1,4,5-trisphosphate-induced Ca2+ release in smooth muscle cells of the guinea pig Taenia caeci. J Gen Physiol. 1990;95:1103–1122. doi: 10.1085/jgp.95.6.1103. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Khodakhah K, Ogden D. Fast activation and inactivation of inositol trisphosphate-evoked Ca2+ release in rat cerebellar Purkinje neurones. J Physiol. 1995;487:343–358. doi: 10.1113/jphysiol.1995.sp020884. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kindman LA, Meyer T. Use of intracellular Ca2+ stores from rat basophilic leukemia cells to study the molecular mechanism leading to quantal Ca2+ release by inositol 1,4,5-trisphosphate. Biochemistry. 1993;32:1270–1277. doi: 10.1021/bi00056a011. [DOI] [PubMed] [Google Scholar]
- Knight PJ, Grigliatti TA. Chimeric G proteins extend the range of insect cell-based functional assays for human G protein-coupled receptors. J Recept Signal Transduct Res. 2004;24:241–256. doi: 10.1081/rrs-200035217. [DOI] [PubMed] [Google Scholar]
- Knight PJ, Pfeifer TA, Grigliatti TA. A functional assay for G-protein-coupled receptors using stably transformed insect tissue culture cell lines. Anal Biochem. 2003;320:88–103. doi: 10.1016/s0003-2697(03)00354-3. [DOI] [PubMed] [Google Scholar]
- Lechleiter J, Girard S, Peralta E, Clapham D. Spiral calcium wave propagation and annihilation in Xenopus laevis oocytes. Science. 1991;252:123–126. doi: 10.1126/science.2011747. [DOI] [PubMed] [Google Scholar]
- Mak D-OD, Foskett JK. Single-channel inositol 1,4,5-trisphosphate receptor currents revealed by patch clamp of isolated Xenopus oocyte nuclei. J Biol Chem. 1994;269:29375–29378. [PubMed] [Google Scholar]
- Mak D-OD, Foskett JK. Single-channel kinetics, inactivation, and spatial distribution of inositol trisphosphate (IP3) receptors in Xenopus oocyte nucleus. J Gen Physiol. 1997;109:571–587. doi: 10.1085/jgp.109.5.571. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mak D-OD, Foskett JK. Effects of divalent cations on single-channel conduction properties of Xenopus IP3 receptor. Am J Physiol. 1998;275:C179–C188. doi: 10.1152/ajpcell.1998.275.1.C179. [DOI] [PubMed] [Google Scholar]
- Mak D-OD, Mcbride S, Foskett JK. Inositol 1,4,5-trisphosphate activation of inositol trisphosphate receptor Ca2+ channel by ligand tuning of Ca2+ inhibition. Proc Natl Acad Sci U S A. 1998;95:15821–15825. doi: 10.1073/pnas.95.26.15821. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mak D-OD, Mcbride S, Foskett JK. Regulation by Ca2+ and inositol 1,4,5-trisphosphate (InsP3) of single recombinant type 3 InsP3 receptor channels. Ca2+ activation uniquely distinguishes types 1 and 3 InsP3 receptors. J Gen Physiol. 2001;117:435–446. doi: 10.1085/jgp.117.5.435. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mak D-OD, Mcbride SM, Foskett JK. Spontaneous channel activity of the inositol 1,4,5-trisphosphate (InsP3) receptor (InsP3R). Application of allosteric modeling to calcium and InsP3 regulation of InsP3R single-channel gating. J Gen Physiol. 2003a;122:583–603. doi: 10.1085/jgp.200308809. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mak D-OD, Mcbride SM, Petrenko NB, Foskett JK. Novel regulation of calcium inhibition of the inositol 1,4,5-trisphosphate receptor calcium-release channel. J Gen Physiol. 2003b;122:569–581. doi: 10.1085/jgp.200308808. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mak D-OD, Mcbride S, Raghuram V, Yue Y, Joseph SK, Foskett JK. Single-channel properties in endoplasmic reticulum membrane of recombinant type 3 inositol trisphosphate receptor. J Gen Physiol. 2000;115:241–256. doi: 10.1085/jgp.115.3.241. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mak D-OD, White C, Ionescu L, Foskett JK. Nuclear patch clamp electrophysiology of inositol trisphosphate receptor Ca2+ release channels. In: Putney JW Jr, editor. Methods in Calcium Signaling Research. 2. Boca Raton: CRC Press; 2005. [Google Scholar]
- Marchant JS, Taylor CW. Cooperative activation of IP3 receptors by sequential binding of IP3 and Ca2+ safeguards against spontaneous activity. Curr Biol. 1997;7:510–518. doi: 10.1016/s0960-9822(06)00222-3. [DOI] [PubMed] [Google Scholar]
- Marchant JS, Taylor CW. Rapid activation and partial inactivation of inositol trisphosphate receptors by inositol trisphosphate. Biochemistry. 1998;37:11524–11533. doi: 10.1021/bi980808k. [DOI] [PubMed] [Google Scholar]
- Marchenko SM, Yarotskyy VV, Kovalenko TN, Kostyuk PG, Thomas RC. Spontaneously active and InsP3-activated ion channels in cell nuclei from rat cerebellar Purkinje and granule neurones. J Physiol. 2005;565:897–910. doi: 10.1113/jphysiol.2004.081299. [DOI] [PMC free article] [PubMed] [Google Scholar]
- McCarron JG, Macmillan D, Bradley KN, Chalmers S, Muir TC. Origin and mechanisms of Ca2+ waves in smooth muscle as revealed by localized photolysis of caged inositol 1,4,5-trisphosphate. J Biol Chem. 2004;279:8417–8427. doi: 10.1074/jbc.M311797200. [DOI] [PubMed] [Google Scholar]
- Meyer T, Stryer L. Transient calcium release induced by successive increments of inositol 1,4,5-trisphosphate. Proc Natl Acad Sci U S A. 1990;87:3841–3845. doi: 10.1073/pnas.87.10.3841. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Missiaen L, Parys JB, De Smedt H, Oike M, Casteels R. Partial calcium release in response to submaximal inositol 1,4,5-trisphosphate receptor activation. Mol Cell Endocrinol. 1994;98:147–156. doi: 10.1016/0303-7207(94)90133-3. [DOI] [PubMed] [Google Scholar]
- Muallem S, Pandol SJ, Beeker TG. Hormone-evoked calcium release from intracellular stores is a quantal process. J Biol Chem. 1989;264:205–212. [PubMed] [Google Scholar]
- Oancea E, Meyer T. Reversible desensitization of inositol trisphosphate-induced calcium release provides a mechanism for repetitive calcium spikes. J Biol Chem. 1996;271:17253–17260. doi: 10.1074/jbc.271.29.17253. [DOI] [PubMed] [Google Scholar]
- Ogden D, Capiod T. Regulation of Ca2+ release by InsP3 in single guinea pig hepatocytes and rat Purkinje neurons. J Gen Physiol. 1997;109:741–756. doi: 10.1085/jgp.109.6.741. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Oldershaw KA, Richardson A, Taylor CW. Prolonged exposure to inositol 1,4,5-trisphosphate does not cause intrinsic desensitization of the intracellular Ca2+-mobilizing receptor. J Biol Chem. 1992;267:16312–16316. [PubMed] [Google Scholar]
- Parker I, Choi J, Yao Y. Elementary events of InsP3-induced Ca2+ liberation in Xenopus oocytes: hot spots, puffs and blips. Cell Calcium. 1996;20:105–121. doi: 10.1016/s0143-4160(96)90100-1. [DOI] [PubMed] [Google Scholar]
- Parker I, Ivorra I. Inhibition by Ca2+ of inositol trisphosphate-mediated Ca2+ liberation: a possible mechanism for oscillatory release of Ca2+ Proc Natl Acad Sci U S A. 1990a;87:260–264. doi: 10.1073/pnas.87.1.260. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Parker I, Ivorra I. Localized all-or-none calcium liberation by inositol trisphosphate. Science. 1990b;250:977–979. doi: 10.1126/science.2237441. [DOI] [PubMed] [Google Scholar]
- Parker I, Yao Y. Calcium puffs in Xenopus oocytes. Ciba Found Symp. 1995;188:50–60. doi: 10.1002/9780470514696.ch4. [DOI] [PubMed] [Google Scholar]
- Parker I, Yao Y. Ca2+ transients associated with openings of inositol trisphosphate-gated channels in Xenopus oocytes. J Physiol. 1996;491:663–668. doi: 10.1113/jphysiol.1996.sp021247. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Parys JB, Missiaen L, De Smedt H, Sienaert I, Henning RH, Casteels R. Quantal release of calcium in permeabilized A7r5 cells is not caused by intrinsic inactivation of the inositol trisphosphate receptor. Biochem Biophys Res Commun. 1995;209:451–456. doi: 10.1006/bbrc.1995.1523. [DOI] [PubMed] [Google Scholar]
- Parys JB, Missiaen L, Smedt HD, Sienaert I, Casteels R. Mechanisms responsible for quantal Ca2+ release from inositol trisphosphate-sensitive calcium stores. Pflugers Arch. 1996;432:359–367. doi: 10.1007/s004240050145. [DOI] [PubMed] [Google Scholar]
- Payne R, Flores TM, Fein A. Feedback inhibition by calcium limits the release of calcium by inositol trisphosphate in Limulus ventral photoreceptors. Neuron. 1990;4:547–555. doi: 10.1016/0896-6273(90)90112-s. [DOI] [PubMed] [Google Scholar]
- Qin F, Auerbach A, Sachs F. A direct optimization approach to hidden Markov modeling for single channel kinetics. Biophys J. 2000;79:1915–1927. doi: 10.1016/S0006-3495(00)76441-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ross SM, Taverna FA, Pickering DS, Wang LY, Macdonald JF, Pennefather PS, Hampson DR. Expression of functional metabotropic and ionotropic glutamate receptors in baculovirus-infected insect cells. Neurosci Lett. 1994;173:139–142. doi: 10.1016/0304-3940(94)90168-6. [DOI] [PubMed] [Google Scholar]
- Sachs F, Qin F, Palade P. Models of Ca2+ release channel adaptation. Science. 1995;267:2010–2011. doi: 10.1126/science.7701327. [DOI] [PubMed] [Google Scholar]
- Sneyd J, Dufour JF. A dynamic model of the type-2 inositol trisphosphate receptor. Proc Natl Acad Sci U S A. 2002;99:2398–2403. doi: 10.1073/pnas.032281999. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sneyd J, Falcke M, Dufour JF, Fox C. A comparison of three models of the inositol trisphosphate receptor. Prog Biophys Mol Biol. 2004;85:121–140. doi: 10.1016/j.pbiomolbio.2004.01.013. [DOI] [PubMed] [Google Scholar]
- Srikanth S, Wang Z, Tu H, Nair S, Mathew MK, Hasan G, Bezprozvanny I. Functional properties of the Drosophila melanogaster inositol 1,4,5-trisphosphate receptor mutants. Biophys J. 2004;86:3634–3646. doi: 10.1529/biophysj.104.040121. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sun XP, Callamaras N, Marchant JS, Parker I. A continuum of InsP3-mediated elementary Ca2+ signalling events in Xenopus oocytes. J Physiol. 1998;509:67–80. doi: 10.1111/j.1469-7793.1998.067bo.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Swatton JE, Morris SA, Wissing F, Taylor CW. Functional properties of Drosophila inositol trisphosphate receptors. Biochem J. 2001;359:435–441. doi: 10.1042/0264-6021:3590435. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Swatton JE, Taylor CW. Fast biphasic regulation of type 3 inositol trisphosphate receptors by cytosolic calcium. J Biol Chem. 2002;277:17571–17579. doi: 10.1074/jbc.M200524200. [DOI] [PubMed] [Google Scholar]
- Taylor CW. Inositol trisphosphate receptors: Ca2+-modulated intracellular Ca2+ channels. Biochim Biophys Acta. 1998;1436:19–33. doi: 10.1016/s0005-2760(98)00122-2. [DOI] [PubMed] [Google Scholar]
- Taylor CW, Potter BV. The size of inositol 1,4,5-trisphosphate-sensitive Ca2+ stores depends on inositol 1,4,5-trisphosphate concentration. Biochem J. 1990;266:189–194. doi: 10.1042/bj2660189. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Thomas D, Lipp P, Berridge MJ, Bootman MD. Hormone-evoked elementary Ca2+ signals are not stereotypic, but reflect activation of different size channel clusters and variable recruitment of channels within a cluster. J Biol Chem. 1998;273:27130–27136. doi: 10.1074/jbc.273.42.27130. [DOI] [PubMed] [Google Scholar]
- Thorn P, Moreton R, Berridge M. Multiple, coordinated Ca2+-release events underlie the inositol trisphosphate-induced local Ca2+ spikes in mouse pancreatic acinar cells. EMBO J. 1996;15:999–1003. [PMC free article] [PubMed] [Google Scholar]
- Tu H, Miyakawa T, Wang Z, Glouchankova L, Iino M, Bezprozvanny I. Functional characterization of the type 1 inositol 1,4,5-trisphosphate receptor coupling domain SII (+/−) splice variants and the opisthotonos mutant form. Biophys J. 2002;82:1995–2004. doi: 10.1016/S0006-3495(02)75548-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tu H, Tang TS, Wang Z, Bezprozvanny I. Association of type 1 inositol 1,4,5-trisphosphate receptor with AKAP9 (Yotiao) and protein kinase A. J Biol Chem. 2004;279:19375–19382. doi: 10.1074/jbc.M313476200. [DOI] [PubMed] [Google Scholar]
- Wilcox RA, Strupish J, Nahorski SR. Quantal calcium release in electropermeabilized SH-SY5Y neuroblastoma cells perfused with myo-inositol 1,4,5-trisphosphate. Cell Calcium. 1996;20:243–255. doi: 10.1016/s0143-4160(96)90030-5. [DOI] [PubMed] [Google Scholar]
- Yao Y, Choi J, Parker I. Quantal puffs of intracellular Ca2+ evoked by inositol trisphosphate in Xenopus oocytes. J Physiol. 1995;482:533–553. doi: 10.1113/jphysiol.1995.sp020538. [DOI] [PMC free article] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.






