Skip to main content
PLOS Computational Biology logoLink to PLOS Computational Biology
. 2014 Aug 14;10(8):e1003783. doi: 10.1371/journal.pcbi.1003783

A Deterministic Model Predicts the Properties of Stochastic Calcium Oscillations in Airway Smooth Muscle Cells

Pengxing Cao 1, Xiahui Tan 2, Graham Donovan 1, Michael J Sanderson 2, James Sneyd 1,*
Editor: Andrew D McCulloch3
PMCID: PMC4133161  PMID: 25121766

Abstract

The inositol trisphosphate receptor (Inline graphic) is one of the most important cellular components responsible for oscillations in the cytoplasmic calcium concentration. Over the past decade, two major questions about the Inline graphic have arisen. Firstly, how best should the Inline graphic be modeled? In other words, what fundamental properties of the Inline graphic allow it to perform its function, and what are their quantitative properties? Secondly, although calcium oscillations are caused by the stochastic opening and closing of small numbers of Inline graphic, is it possible for a deterministic model to be a reliable predictor of calcium behavior? Here, we answer these two questions, using airway smooth muscle cells (ASMC) as a specific example. Firstly, we show that periodic calcium waves in ASMC, as well as the statistics of calcium puffs in other cell types, can be quantitatively reproduced by a two-state model of the Inline graphic, and thus the behavior of the Inline graphic is essentially determined by its modal structure. The structure within each mode is irrelevant for function. Secondly, we show that, although calcium waves in ASMC are generated by a stochastic mechanism, Inline graphic stochasticity is not essential for a qualitative prediction of how oscillation frequency depends on model parameters, and thus deterministic Inline graphic models demonstrate the same level of predictive capability as do stochastic models. We conclude that, firstly, calcium dynamics can be accurately modeled using simplified Inline graphic models, and, secondly, to obtain qualitative predictions of how oscillation frequency depends on parameters it is sufficient to use a deterministic model.

Author Summary

The inositol trisphosphate receptor (Inline graphic) is one of the most important cellular components responsible for calcium oscillations. Over the past decade, two major questions about the Inline graphic have arisen. Firstly, what fundamental properties of the Inline graphic allow it to perform its function? Secondly, although calcium oscillations are caused by the stochastic properties of small numbers of Inline graphic is it possible for a deterministic model to be a reliable predictor of calcium dynamics? Using airway smooth muscle cells as an example, we show that calcium dynamics can be accurately modeled using simplified Inline graphic models, and, secondly, that deterministic models are qualitatively accurate predictors of calcium dynamics. These results are important for the study of calcium dynamics in many cell types.

Introduction

Oscillations in cytoplasmic calcium concentration (Inline graphic), mediated by inositol trisphosphate receptors (Inline graphic; a calcium channel that releases calcium ions (Inline graphic) from the endoplasmic or sarcoplasmic reticulum (ER or SR) in the presence of inositol trisphosphate (Inline graphic)) play an important role in cellular function in many cell types. Hence, a thorough knowledge of the behavior of the Inline graphic is a necessary prerequisite for an understanding of intracellular Inline graphic oscillations and waves. Mathematical and computational models of the Inline graphic play a vital role in studies of Inline graphic dynamics. However, over the past decade, two major questions about Inline graphic models have arisen.

Firstly, how best should the Inline graphic be modeled? Models of the Inline graphic have a long history, beginning with the heuristic models of [1][3]. With the recent appearance of single-channel data from Inline graphic in vivo [4], [5], a new generation of Markov Inline graphic models has recently appeared [6], [7]. These models show that Inline graphic exist in different modes with different open probabilities. Within each mode there are multiple states, some open, some closed. Importantly, it was found [8] that time-dependent transitions between different modes are crucial for reproducing Inline graphic puff data from [9]. However, it is not yet clear whether transitions between states within each mode are important, or whether all the important behaviors are captured simply by inter-mode transitions.

Secondly, why do deterministic models of the Inline graphic perform so well as predictive models? Deterministic models of the Inline graphic have proven to be useful predictive models in a range of cell types. For example, Inline graphic-based models have been developed to study Inline graphic oscillations in airway smooth muscle cells (ASMC) [10][13], and these models have made predictions which have been confirmed experimentally. This shows the usefulness of such models in advancing our understanding of how intracellular Inline graphic oscillations and waves are initiated and controlled in ASMC. However, these models are deterministic models which assume infinitely many Inline graphic per unit cell volume, an assumption that contradicts experimental findings in many cell types showing that Inline graphic puffs and spikes occur stochastically, and that intracellular Inline graphic waves and oscillations arise as an emergent property of fundamental stochastic events [9], [14], [15].

Here, we answer these two fundamental modeling questions using data and models from ASMC. Firstly, we show that a simple model of the Inline graphic, involving only two states with time-dependent transitions, suffices to generate correct dynamics of Inline graphic puffs and oscillations. Secondly, we show that, although Inline graphic oscillations in ASMC are generated by a stochastic mechanism, a deterministic model can make the same qualitative predictions as the analogous stochastic model, indicating that deterministic models, that require much less computational time and complexity, can be used to make reliable predictions. Although we work in the specific context of ASMC, our results are applicable to other cell types that exhibit similar Inline graphic oscillations and waves.

Results

A two-state model of the Inline graphic is sufficient to reproduce function

We have previously shown [8] that the statistics of Inline graphic puffs in SH-SY5Y cells can be reproduced by a Markov model of the Inline graphic based on the steady-state data of [5] and the time-dependent data of [4]. In this model the Inline graphic can exist in 6 different states, grouped into two modes, which we call Drive and Park (see Fig. 1). The Drive mode (which contains 4 states; 1 open and 3 closed) has an average open probability of around 0.7, while the Park mode (which contains the remaining two states; 1 open and 1 closed) has an open probability close to zero. Transitions between states within each mode are independent of Inline graphic and Inline graphic; only the transitions between modes are ligand-dependent.

Figure 1. The structure of the Siekmann Inline graphic model.

Figure 1

The Inline graphic model is comprised of two modes. One is the drive mode containing three closed states Inline graphic, Inline graphic, Inline graphic and one open state Inline graphic. The other is the park mode which includes one closed state Inline graphic and one open state Inline graphic. Inline graphic are rates of state-transitions between two adjacent states and Inline graphic and Inline graphic are transitions between the two modes [7].

In our previous study on calcium puffs [8], we showed that, to reproduce the experimentally observed non-exponential interspike interval (ISI) distribution and coefficient of variation (CV) of ISI smaller than 1, the time-dependent intermodal transitions are crucial. Lack of time dependencies in the Siekmann model leads to exponential ISI distributions and CV = 1, which is not the case for calcium spikes in ASMC. Fig. 2A shows an example of Inline graphic oscillations generated by 50 nM methacholine (MCh, an agonist that can induce the production of Inline graphic by binding to a G protein-coupled receptor in the cell membrane) in ASMC. By gathering data from 14 cells in 5 mouse lung slices, we found that the standard deviation of the interspike interval (ISI) is approximately a linear function of the ISI mean, with a slope clearly between 0 and 1 (i.e. Inline graphic), indicating that the spikes are generated by an inhomogeneous Poisson process (a slope of 1 would denote a pure Poisson process) (see Fig. 2B). This shows the necessity of inclusion of time-dependent transitions for mode-switching.

Figure 2. Inline graphic oscillations in ASMC in lung slices are generated by a stochastic mechanism.

Figure 2

A: experimental Inline graphic spiking in ASMC in lung slices, stimulated with 50 nM MCh. In the upper panel we filter out baseline noise by using a low threshold of 1.42 (relative fluorescence intensity) and then choose samples with amplitude larger than 1.75. The ISI calculated from the upper panel is shown in the lower panel. B: relationship between the standard deviation and the mean of experimental ISIs. Data obtained from 14 ASMC in 5 mouse lung slices. The relationship is approximately linear with a slope of 0.66, which implies that an inhomogeneous Poisson process governs the generation of oscillations. The dashed line indicates where the coefficient of variation (CV) is 1 (as it is for a pure Poisson process). Variation in ISI is mainly caused by both use of different doses of MCh and different sensitivities of different cells to MCh. Error bars indicate the standard errors of the means (SEM).

Using a quasi-steady-state approximation, and ignoring states with very low dwell times, it is possible to construct a simplified two-state version of the full six-state model (see Materials and Methods ). In the simplified model the intramodal structure is ignored, and only the intermodal transitions have an effect on Inline graphic behavior. In Fig. 3 we compared the simplified Inline graphic model to the full six-state model. Both models have the same distribution of interspike interval, spike amplitude and spike duration. Moreover, by looking at a more detailed comparison between the two model results (Figs. 4A, C and E) and experimental data (Figs. 4B, D and F), we found the 2-state model not only can reproduce the behaviour of the 6-state model, but can also qualitatively reproduce experimental data. The average experimental ISI shows a clear decreasing trend as MCh concentration increases (although a saturation occurs in the data for high MCh), a trend that is mirrored by the model results as the Inline graphic concentration increases. Unfortunately, since the exact relationship between MCh concentration and Inline graphic concentration is uncertain, a quantitative comparison is not possible. In both model and experimental results, the average peak and duration of the oscillations are nearly independent of agonist concentration. The quantitative difference in spike duration between the model results and the data in Figs. 4E and F are most likely due to choice of calcium buffering parameters. For example, adding Inline graphic fast Inline graphic buffer (see Materials and Methods ) increases the average spike duration to 0.54 s or 0.7 s respectively, which are close to the levels shown in the data.

Figure 3. A 2-state open/closed model quantitatively reproduces the 6-state Inline graphic model.

Figure 3

A: histograms of interspike interval (ISI) distribution for both the 6-state and the simplified models. The ISI is defined to be the waiting time between successive spikes. Each histogram contain an equal number of samples (180). B: comparison of average ISI, average peak value of Inline graphic (Inline graphic in the model) and average spike duration. All distributions were computed at a constant Inline graphic.

Figure 4. More detailed comparisons between the 2-state and the 6-state Inline graphic models, and a comparison to experimental data.

Figure 4

As a function of Inline graphic concentration (Inline graphic), the two models give the same ISI (A), peak Inline graphic (C) and spike duration (E). These results agree qualitatively with experimental data, as shown in panels B, D and F respectively. Quantitative comparisons are generally not possible as the relationship between Inline graphic concentration and agonist concentration is not known. Error bars represent Inline graphic. Data for each MCh concentration are obtained from at least three different cells from at least two different lung slices.

Thus, the intramodal structure of the six-state model is essentially unimportant, as the model behavior (in terms of the statistics of puffs and oscillations) is governed almost entirely by the time dependence of the intermode transitions, particularly the time dependence of the rapid inhibition of the Inline graphic by high Inline graphic, and the slow recovery from inhibition by Inline graphic. The multiple states within each mode are necessary to obtain an acceptable quantitative fit to single-channel data, but are nevertheless of limited importance for function. Hence, even when simulating microscopic events such as Inline graphic puffs it is sufficient to use a simpler, faster, two-state model, rather than a more complex six-state model. In the following, we will use the 2-state Inline graphic model to generate all the simulation results.

Prediction of stochastic Inline graphic behavior by a deterministic model

Although the data (Fig. 2) show that Inline graphic oscillations in ASMC are generated by a stochastic process, not a deterministic one, we wish to know to what extent a deterministic model can be used to make qualitative (and experimentally testable) predictions. Our simplified 2-state Markov model of the Inline graphic can be converted to a deterministic model (see Materials and Methods ). The result is a system of ordinary differential equations (ODEs) with four variables, which takes into account the increased Inline graphic at an open Inline graphic pore, as well as the increased Inline graphic within a cluster of Inline graphic; the four variables are the Inline graphic outside the Inline graphic cluster (Inline graphic), the Inline graphic within the Inline graphic cluster (Inline graphic), the total intracellular Inline graphic concentration (Inline graphic) and an Inline graphic gating variable (Inline graphic). We refer to the reduced 4D model as the deterministic model for all the results and analyses.

Note that there is no physical or geometric constraint enforcing a high local Inline graphic; in this case the spatial heterogeneity arises solely from the low diffusion coefficient of Inline graphic. Our use of Inline graphic is merely a highly simplified way of introducing spatial heterogeneity of the Inline graphic concentration. Since the Inline graphic can only “see” Inline graphic (as well as the Inline graphic concentration right at the mouth of an open channel, which we denote by Inline graphic), but cannot be influenced directly by Inline graphic (the experimentally observed Inline graphic signal), our approach allows for the functional differentiation of the rapid local oscillatory Inline graphic in the cluster, from the slower Inline graphic signal in the cytoplasm, without the need for computationally intensive simulations of a partial differential equation model. Quantitative accuracy is thus sacrificed for computational convenience.

Calcium oscillations in the stochastic and deterministic models are shown in Fig. 5A. According to our previous results [8], the average value of Inline graphic over the cluster of Inline graphic primarily regulates the termination and regeneration of individual spikes. This can be seen in the stochastic model by projecting the solution on the Inline graphic phase plane (Fig. 5B). Upon an initial Inline graphic release from one or more Inline graphic, a large spike is generated by Ca2+-induced Inline graphic release (via the Inline graphic) during which time a decreasing Inline graphic gradually decreases the average open probability of the clustered Inline graphic. The spike is terminated when Inline graphic is too small to allow further Inline graphic release. This phenomenon is qualitatively reproduced by the deterministic model (Fig. 5D). In both the stochastic and deterministic models the decrease in average Inline graphic open probability of a cluster of Inline graphic caused by Inline graphic inhibition is the main reason for the termination of each spike.

Figure 5. Stochastic and deterministic simulations exhibit similar dynamic properties.

Figure 5

A: simulated stochastic (upper panel) or deterministic (lower panel) Inline graphic oscillations at Inline graphic Inline graphic. B: a typical stochastic solution projected on the Inline graphic plane. The average Inline graphic represents the average value of Inline graphic over the 20 Inline graphic. Statistics (Inline graphic) of the initiation point (blue square), the peak (red square) and termination point (green square) are shown in the inset. 116 samples are obtained by applying a low threshold of Inline graphic and a high threshold of Inline graphic to Inline graphic. C: a typical periodic solution of the deterministic model (black curve), plotted in the Inline graphic phase space. The arrow indicates the direction of movement. Inline graphic is the slowest variable so that its variation during an oscillation is very small. This allows to treat Inline graphic as a constant (Inline graphic in this case) and study the dynamics of the model in the Inline graphic phase space. The color surface is the surface where Inline graphic (called the critical manifold). The white N-shaped curve is the intersection of the critical manifold and the surface Inline graphic. D: projection of the periodic solution to the Inline graphic plane. The red N-shaped curve is the projection to the Inline graphic plane of the white curve shown in C. The evolution of the deterministic solution exhibits three different time scales separated by green circles (labelled by a, b and c) and indicated by arrows (triple arrow: fastest; double arrow: intermediate; single arrow: slowest).

According to Figs. 5B and D, regeneration of each spike requires a return of Inline graphic back to a relatively high value (i.e., recovery of the Inline graphic from inhibition by Inline graphic). The deterministic model sets a clear threshold for the regeneration, as can be seen in Fig. 5C, where an upstroke in Inline graphic occurs when the trajectory creeps beyond the sharp “knee” of the white curve. When the trajectory reaches the knees of the white curve it is forced to jump across to the other stable branch of the critical manifold, resulting in a fast increase in Inline graphic followed by a relatively fast increase in Inline graphic (seen by combining Figs. 5C and D).

In contrast, the stochastic model enlarges the contributions of individual Inline graphic so that the generation of each spike is also effectively driven by random Inline graphic release through the Inline graphic, which can be seen in the inset of Fig. 5B where the site of spike initiation (blue bar) exhibits significantly greater variation than that of spike termination (green bar). In spite of this, the essential similarities in phase plane behavior result in both deterministic and stochastic models making the same qualitative predictions in response to perturbations, such as changes in Inline graphic concentration (Inline graphic), Inline graphic influx or efflux. In the following, we illustrate this by investigating a number of experimentally testable predictions. Due to the extensive importance of frequency encoding in many Inline graphic-dependent processes, we focus particularly on the change of oscillation frequency in response to parameter perturbations. As a side issue we also investigate how the oscillation baseline depends on physiologically important parameters.

Dependence of oscillation frequency on Inline graphic concentration

In many cell types a moderate increase in Inline graphic increases the Inline graphic oscillation frequency (see Fig. 2A in [11], Fig. 4E in [16] and Fig. 6B in [17]), a result that is reproduced by both model types (Fig. 6A). As Inline graphic increases, the stochastic model increases the probability of the initial Inline graphic release through the first open Inline graphic and of the following Inline graphic release, thus shortening the average ISI. Although the oscillatory region of the deterministic model is strictly confined by bifurcations which do not apply to the stochastic model, the deterministic model can successfully replicate an increasing frequency by lowering the “knee” of the red curve in Fig. 5D and shortening the time spent from the termination point c to the initiation point a (thus shortening the ISI). Hence, although the deterministic model cannot be used to predict the exact values of Inline graphic at which the oscillations begin and end, as stochastic effects predominate in these regions, it can be used to predict the correct qualitative trend in oscillation frequency.

Figure 6. Comparison of parameter-dependent frequency changes in the stochastic and deterministic models.

Figure 6

All curves are computed at Inline graphic Inline graphic except in panel A, which uses a variety of Inline graphic. Other parameters are set at their default values given in Table 1. A: as Inline graphic increases, Inline graphic oscillations in both models increase in frequency. B: as Inline graphic influx increases (modeled by an increase in receptor-operated calcium channel flux coefficient Inline graphic), so does the oscillation frequency in both models. C: as Inline graphic efflux increases (modeled by an increase in plasma pump expression Inline graphic), oscillation frequency decreases. D: as SERCA pump expression, Inline graphic, increases, so does oscillation frequency. E: as total buffer concentration, Inline graphic, increases, oscillation frequency decreases.

Dependence of oscillation frequency on Inline graphic influx and efflux

In many cell types, including ASMC, transmembrane fluxes modulate the total intracellular Inline graphic load (Inline graphic) on a slow time scale [16], [18], and thereby modulate the oscillation frequency [19]. Experimental data can be seen in Fig. 8 in [16] and Fig. 2 in [18]. Figs. 6B and C show that both stochastic and deterministic models predict the same qualitative changes in oscillation frequency in response to changes in membrane fluxes (through membrane ATPase pumps and/or Inline graphic influx channels such as receptor-operated channels or store-operated channels).

Figure 8. Schematic diagram of the Inline graphic model.

Figure 8

Inline graphic represents cytoplasmic Inline graphic concentration, excluding a small local Inline graphic (whose concentration is denoted by Inline graphic) close to the Inline graphic release site (i.e., an Inline graphic cluster). Upon coordinated openings of the Inline graphic, SR Inline graphic (Inline graphic) is first released into the local domain (Inline graphic) to cause a rapid increase in Inline graphic. High local Inline graphic then diffuses to the rest of the cytoplasm (Inline graphic), and is eventually pumped back to the SR (Inline graphic).

Dependence of oscillation frequency on SERCA expression

The level of sarco/endoplasmic reticulum calcium ATPase (SERCA) expression (or capacity) is important for airway remodeling in asthma [20] and ASMC Inline graphic oscillations [21]. We thus investigated the predictions of the two models in response to changes in SERCA expression (Inline graphic). As Inline graphic decreases, the deterministic model exhibits a decreasing frequency, in agreement with experimental data (see Figs. 3 and 4 in [21]). The same trend is seen in the stochastic model with only 20 Inline graphic (see Fig. 6D).

Dependence of oscillation frequency on Inline graphic buffer concentration

Calcium buffers have been shown to be able to change the ISI and spike duration, which in turn change the oscillation frequency [15], [22]. We compared the effects on the two models of varying total buffer concentration (Inline graphic) by adding one buffer with relatively fast kinetics to the models (see Materials and Methods for details). In both models the frequency decreases as Inline graphic increases (see Fig. 6E), which is consistent with experimental data (Fig. 2B in [18]). This is not surprising, because increasing Inline graphic can decrease the effective rates of SR Inline graphic release and reuptake.

Dependence of oscillation baseline on Inline graphic influx and SERCA expression

Sustained elevations of baseline during agonist-induced Inline graphic oscillations or transients have been observed experimentally, and are believed to be a result of an increase in Inline graphic influx caused by opening of membrane Inline graphic channels [13], [16]. Furthermore, there is evidence showing that decreased SERCA expression could also increase the baseline (Fig. 4 in [21]). Those phenomena are successfully reproduced by both models (see Fig. 7).

Figure 7. Dependence of calcium oscillation baseline on calcium influx and SERCA expression.

Figure 7

A: increasing influx (described by Inline graphic) increases the average trough of Inline graphic oscillations. B: decreasing SERCA expression (described by Inline graphic) increases the average trough of Inline graphic oscillations. All curves are computed at Inline graphic Inline graphic.

Discussion

In this paper we address two current major questions in the field of Inline graphic modeling. Firstly, we show that Inline graphic puffs and stochastic oscillations can be reproduced quantitatively by an extremely simple model, consisting only of two states (one open, one closed), with time-dependent transitions between them. This model is obtained by removing the intramodal structure of a more complex model that was determined by fitting a Markov model to single-channel data [7]. We thus show that the internal structure of each mode is irrelevant for function and mode switching is the key mechanism for the control of calcium release. The necessity for time-dependent mode switching is shown not only by the dynamic single-channel data of [4]), but also by the puff data of [9] and our ASMC data.

Secondly, we investigate the role of stochasticity of Inline graphic in modeling Inline graphic oscillations in ASMC by comparing a stochastic IP3R-based Inline graphic model and its associated deterministic version, for parameters such that both of the models exhibit Inline graphic spikes but the stochastic model cannot necessarily be replaced by a mean-field model. We find that a four-variable deterministic model has the same predictive power as the stochastic model, in that it correctly reproduces the process of spike termination and predicts the same qualitative changes in oscillation frequency and baseline in response to a variety of perturbations that are commonly used experimentally. The mechanism for termination of individual spikes is fundamentally a deterministic process controlled by a rapid inhibition induced by the high local Inline graphic in the Inline graphic cluster, whereas spike initiation is significantly affected by stochastic opening of Inline graphic. Hence, repetitive Inline graphic cycling is primarily induced by the time-dependent gating variables governing transitions of the Inline graphic from one mode to another.

Our simplified two-state model of the Inline graphic is identical in structure (although not in parameter values) to the well-known model of [23]. It is somewhat ironic that after 20 years of detailed studies of the Inline graphic and the construction of a plethora of models of varying complexity, the single-channel data have led us around full circle, back to these original formulations. Excitability is arising via a fast activation followed by a slower inactivation, a combination often seen in physiological processes [24]. Encoding of this fundamental combination results directly from the two-mode structure of the Inline graphic. Although similar single-channel data have been used to construct three-mode models [6], [25], neither of these models has yet been used in detailed studies of Inline graphic puffs and waves, and it remains unclear whether or not they have a similar underlying structure.

In contrast to previous deterministic ODE models, our four-variable Inline graphic model includes a more accurate Inline graphic model, as well as local control of clustered Inline graphic by two distinct Inline graphic microdomains; one at the mouth of an open Inline graphic, the other inside a cluster of Inline graphic. Neglect of either of these microdomains leads to models that either exhibit unphysiological cytoplasmic Inline graphic concentrations or fail to reproduce reasonable oscillations. This underlines the importance of taking Inline graphic microdomains into consideration when constructing any model. Our microdomain model is highly simplified, with the microdomain being treated simply as a well-mixed compartment. More detailed modeling of spatially-dependent microdomains is possible, and not difficult in principle, but requires far greater computational resources. It is undeniable that a more detailed model, incorporating the full spatial complexity – and possibly stochastic aspects as well – would make, overall, a better predictive tool. However, our goal is to find the simplest models that can be used as predictive tools.

An important similar study is that of Shuai and Jung [26]. They compared the use of Markov and Langevin approaches to the computation of puff amplitude distributions, compared their results with the deterministic limit, and showed that Inline graphic stochasticity does not qualitatively change the type of puff amplitude distribution except for when there are fewer than 10 Inline graphic. Here, we significantly extend the scope of their study by exploring the effects of Inline graphic stochasticity on the dynamics of Inline graphic spikes, and we do this in the context of an Inline graphic model that has been fitted to single-channel data. Although this is true in a general sense for the Li-Rinzel model, which is based on the DeYoung-Keizer model, which did take into account the opening time distributions of Inline graphic in lipid bilayers, neither model can reproduce the more recent data obtained from on-nuclei patch clamping. When these recent data are taken into account one obtains a model with the same structure, but quite different parameters and behavior.

We find that, in spite of a relatively large variation in spike amplitude which is partially caused by a large variation in ISI (Fig. 5B), the mechanism governing individual spike terminations is the same for both a few or infinitely many Inline graphic, which explains why the one-peak type of amplitude distribution is independent of the choice of Inline graphic number (see Fig. 6A in [26]).

Another important relevant study was done by Dupont et al. [27], who compared the regularity of stochastic oscillations in hepatocytes for different numbers of Inline graphic clusters. They found that the impact of Inline graphic stochasticity on global Inline graphic oscillations (in terms of CV) increases as the total cluster number decreases. Our study here extends these results, and demonstrates how well stochastic oscillations can be qualitatively described by a deterministic system, even when there is only a small number of Inline graphic (which appears to be the case for ASMC, in which the wave initiation site is only Inline graphic in diameter). Indeed, as we have shown, for the purposes of predictive modeling a simple deterministic model does as well as more complex stochastic simulations.

Ryanodine receptors (RyR) are another important component modulating ASMC Inline graphic oscillations [16], [28], [29] but are not included in our model. This is because the role of RyR is not fully understood and may be species-dependent; for example, in mouse or human ASMC, RyR play very little role in Inline graphic-induced continuing Inline graphic oscillations [17], [30], but this appears not to be true for pigs [28]. Our study focuses on the calcium oscillations in mouse and human (as we did in our experiments) where inclusion of a deterministic model of RyR should have little effect. An understanding of the role of RyR stochasticity and how the Inline graphic and the RyR interact needs a reliable RyR Markov model, exclusive to ASMC, which is not currently available. Multiple Markov models of the RyR have been developed for use in cardiac cells [31], but these are based on single-channel data from lipid bilayers, and are adapted for the specific context of cardiac cells. Their applicability to ASMC remains unclear.

Although we have not shown that the deterministic model for ASMC has the same predictive power as the stochastic model in all possible cases (which would hardly be possible in the absence of an analytical proof) the underlying similarity in phase plane structure indicates that such similarity is plausible at least. Certainly, we have not found any counterexample to this claim. However, whether or not this claim is true for all cell types is unclear. Some cell types exhibit both local Inline graphic puffs and global Inline graphic spikes (usually propagating throughout the cells in the form of traveling waves), showing that initiation of such Inline graphic spikes requires a synchronization of Inline graphic release from more than one cluster of Inline graphic [14]. This type of spiking relies on the hierarchical organization of Inline graphic signal pathways, in particular the stochastic recruitment of both individual Inline graphic and puffs at different levels [32], and therefore cannot be simply reproduced by deterministic models containing only a few ODEs. However, Inline graphic oscillations in ASMC, as observed in lung slices, may not be of this type, as IP3R-dependent puffs have not been seen in these ASMC. It thus appears that, in ASMC in lung slices, every Inline graphic “puff” initiates a wave, resulting in periodic waves with ISI that are governed by the dynamics of individual puffs.

Materials and Methods

Ethics Statement

Animal experimentations carried out were approved by the Animal Care and Use Committee of the University of Massachusetts Medical School under approval number A-836-12.

Lung slice preparation

BALB/c mice (7–10 weeks old, Charles River Breeding Labs, Needham, MA) were euthanized via intraperitoneal injection of 0.3 ml sodium pentabarbitone (Oak Pharmaceuticals, Lake Forest, IL). After removal of the chest wall, lungs were inflated with Inline graphic of 1.8% warm agarose in sHBSS via an intratracheal catheter. Subsequently, air (Inline graphic) was injected to push the agarose within the airways into the alveoli. The agarose was polymerized by cooling to Inline graphic. A vibratome (VF-300, Precisionary Instruments, San Jose, CA) was used to make Inline graphic thick slices which were maintained in Dulbecco's Modified Eagle's Media (DMEM, Invitrogen, Carlsbad, CA) at Inline graphic in Inline graphic/air. All experiments were conducted at Inline graphic in a custom-made temperature-controlled Plexiglas chamber as described in [17].

Measurement of Inline graphic oscillations

Lung slices were incubated in sHBSS containing Inline graphic Oregon Green 488 BAPTA-1-AM (Invitrogen), a Ca2+-indicator dye, 0.1% Pluronic F-127 (Invitrogen) and Inline graphic sulfobromophthalein (Sigma Aldrich, St Louis, MO) in the dark at Inline graphic for 1 hour. Subsequently, the slices were incubated in Inline graphic sulfobromophthalein for 30 minutes. Slices were mounted on a cover-glass and held down with Inline graphic mesh. A smaller cover-glass was placed on top of the mesh and sealed at the sides with silicone grease to facilitate solution exchange. Slices were examined with a custom-built 2-photon scanning laser microscope with a Inline graphic oil immersion objective lens and images recorded at 30 images per second using Videosavant 4.0 software (IO Industries, Montreal, Canada). Changes in fluorescence intensity (which represent changes in Inline graphic) were analyzed in an ASMC of interest by averaging the grey value of a Inline graphic pixel region using custom written software. Relative fluorescence intensity (Inline graphic) was expressed as a ratio of the fluorescence intensity at a particular time (F) normalized to the initial fluorescence intensity (Inline graphic).

The calcium model

Inhomogeneity of cytoplasmic Inline graphic concentration not only exists around individual channel pores of the Inline graphic, where a nearly instantaneous high Inline graphic concentration at the pore (denoted by Inline graphic) leads to a very sharp concentration profile, but is also seen inside an Inline graphic cluster where the average cluster Inline graphic concentration (Inline graphic) is apparently higher than that of the surrounding cytoplasm (Inline graphic) [33]. This indicates that during Inline graphic oscillations each Inline graphic is controlled by either the pore Inline graphic concentration (when it is open) or the cluster Inline graphic concentration (when it is closed). Neither of these local concentrations influence cell membrane fluxes or the majority of SERCAs, which we assume to be distributed outside the cluster.

The scale separation between the pore Inline graphic concentration and the cluster Inline graphic concentration allows to treat Inline graphic as a parameter, providing a simpler way of modeling local Inline graphic events (like Inline graphic puffs) that has been used in several previous studies [8], [34], [35]. However, evolution of the cluster concentration and wide-field cytoplasm Inline graphic concentration are not always separable, so an additional differential equation for the cluster Inline graphic is necessary.

A schematic diagram of the model is shown in Fig. 8. The corresponding ODEs are

graphic file with name pcbi.1003783.e341.jpg (1)
graphic file with name pcbi.1003783.e342.jpg (2)
graphic file with name pcbi.1003783.e343.jpg (3)

where Inline graphic representing total intracellular Inline graphic concentration, and thus SR Inline graphic concentration, Inline graphic is given by Inline graphic. Inline graphic and Inline graphic are the volume ratios given in Table 1. Inline graphic is the flux through the Inline graphic, Inline graphic is a background Inline graphic leak out of the SR, and Inline graphic is the uptake of Inline graphic into the SR by SERCA pumps. Inline graphic is the flux through plasma pump, and Inline graphic represents a sum of main Inline graphic influxes including Inline graphic (receptor-operated Inline graphic channel), Inline graphic (store-operated Inline graphic channel) and Inline graphic (Inline graphic leak into the cell). Inline graphic coarsely models the diffusion flux from cluster microdomain to the cytoplasm. Details of the fluxes are

Table 1. Parameter values of the stochastic calcium model.

Parameter Description Value/Units
Inline graphic Inline graphic flux coefficient Inline graphic
Inline graphic Inline graphic diffusional flux coefficient Inline graphic
Inline graphic SR leak flux coefficient Inline graphic
Inline graphic maximum capacity of SERCA Inline graphic
Inline graphic SERCA half-maximal activating Inline graphic Inline graphic
Inline graphic Hill coefficient for SERCA 1.75
Inline graphic plasma membrane leak influx Inline graphic
Inline graphic ROCC flux coefficient Inline graphic
Inline graphic maximum capacity of SOCC Inline graphic
Inline graphic SOCC dissociation constant Inline graphic
Inline graphic maximum capacity of plasma pump Inline graphic
Inline graphic half-maximal activating Inline graphic of plasma pump Inline graphic
Inline graphic Hill coefficient for plasma pump 2
Inline graphic the cytoplasmic-to-microdomain volume ratio 100
Inline graphic the cytoplasmic-to-SR volume ratio 10
Inline graphic an instantaneous high Inline graphic at open channel pore when Inline graphic Inline graphic
Inline graphic total number of Inline graphic channels 20
  • • Different formulations of Inline graphic give different types of models:

    • a) For the stochastic model, Inline graphic where Inline graphic is the maximum conductance of a cluster of Inline graphic Inline graphic (here Inline graphic). Inline graphic is the number of open Inline graphic determined by the states of Inline graphic.

    • b) For the deterministic model we set Inline graphic where Inline graphic is the Inline graphic open probability, a continuous analogue of Inline graphic.

    To calculate Inline graphic and Inline graphic, we use the Inline graphic model of [7], [8], with minor modifications described later.

  • Inline graphic

  • Inline graphic where Inline graphic and Inline graphic are obtained from [36].

  • Inline graphic

  • Inline graphic includes a basal leak (Inline graphic), receptor-operated calcium channel (ROCC, Inline graphic), store-operated calcium channel (SOCC, Inline graphic). By using the Inline graphic concentration (Inline graphic) as a surrogate indicator of MCh concentration, we assume that Inline graphic. SOCC is modeled by Inline graphic [13].

  • Inline graphic

Calcium concentration at open channel pore (Inline graphic) does not explicitly appear in the equations but is used in the Inline graphic model introduced later. Inline graphic is assumed to be proportional to SR Inline graphic concentration (Inline graphic) and is therefore simply modeled by Inline graphic where Inline graphic is the value corresponding to Inline graphic. Alternatively, Inline graphic can also be assumed to be a large constant (say greater than Inline graphic) without fundamentally altering the model dynamics. The choice of Inline graphic is not critical as long as it is sufficiently large to play a role in inactivating the open channels. All the parameter values are given in Table 1.

The data-driven Inline graphic model

The Inline graphic model used in our ASMC calcium model is an improved version of the Siekmann Inline graphic model which is a 6-state Markov model derived by fitting to the stationary single channel data using Markov chain Monte Carlo (MCMC) [5], [7], [8]. Fig. 1 has shown the structure of the Inline graphic model which is comprised of two modes; the drive mode, containing three closed states Inline graphic, Inline graphic, Inline graphic and one open state Inline graphic, and the park mode, containing one closed state Inline graphic and one open state Inline graphic. The transition rates in each mode are constants (shown in Table 2), but Inline graphic and Inline graphic which connect the two modes are Inline graphic-/Inline graphic-dependent and are formulated as

graphic file with name pcbi.1003783.e422.jpg (4)
graphic file with name pcbi.1003783.e423.jpg (5)

where Inline graphic, Inline graphic, Inline graphic and Inline graphic are Inline graphic-/Inline graphic-modulated gating variables. Inline graphic, Inline graphic, Inline graphic and Inline graphic are either functions of Inline graphic or constants and are given later. We assume the gating variables obey the following differential equation,

graphic file with name pcbi.1003783.e435.jpg (6)

where Inline graphic is the equilibrium and Inline graphic is the rate at which the equilibrium is approached. Those equilibria are functions of Inline graphic concentration at the cytoplasmic side of the Inline graphic, denoted by Inline graphic in the equations, equal to either Inline graphic or Inline graphic depending on the state of the channel). They are assumed to be

graphic file with name pcbi.1003783.e443.jpg (7)
graphic file with name pcbi.1003783.e444.jpg (8)
graphic file with name pcbi.1003783.e445.jpg (9)
graphic file with name pcbi.1003783.e446.jpg (10)

Table 2. Parameter values of the Inline graphic model.

Parameter Value/Units Parameter Value/Units
Inline graphic Inline graphic Inline graphic Inline graphic
Inline graphic Inline graphic Inline graphic Inline graphic
Inline graphic Inline graphic Inline graphic Inline graphic
Inline graphic Inline graphic Inline graphic Inline graphic
Inline graphic Inline graphic Inline graphic Inline graphic
Inline graphic Inline graphic Inline graphic Inline graphic
Inline graphic Inline graphic

Hence, we have stationary expressions of Inline graphic and Inline graphic,

graphic file with name pcbi.1003783.e476.jpg (11)
graphic file with name pcbi.1003783.e477.jpg (12)

The expressions of Inline graphics, Inline graphics, Inline graphics and Inline graphics are chosen as follows so that Eq. 11 and Eq. 12 capture the correct trends of experimental values of Inline graphic and Inline graphic (see Fig. 9) and generate relatively smooth open probability curves (see Fig. 10),

graphic file with name pcbi.1003783.e484.jpg

Figure 9. Stationary data and fits of Inline graphic and Inline graphic.

Figure 9

Stationary transition rates of Inline graphic and Inline graphic, Inline graphic and Inline graphic, as functions of Inline graphic concentration were estimated and fitted for two Inline graphic, Inline graphic (A) and Inline graphic (B). Circles and squares represent the means of Inline graphic and Inline graphic distributions computed by MCMC simulation [7]. Note that MCMC failed to determine the values of Inline graphic and Inline graphic at Inline graphic for Inline graphic Inline graphic, as the Inline graphic was almost in the drive mode for these cases. The corresponding fitting curves (solid for Inline graphic; dashed for Inline graphic) are produced using Eqs. 7–12.

Figure 10. Open probability curves for various Inline graphic.

Figure 10

Inline graphic is equal to the sum of probabilities of the Inline graphic in Inline graphic and Inline graphic. Three representative curves correspond to Inline graphic, Inline graphic and Inline graphic Inline graphic (from bottom to top) respectively. Data (average open probability) are from [5].

Note that the above formulas are different from the relatively complicated formulas used in [8]. The rates, Inline graphic, Inline graphic and Inline graphic, are constants estimated by using dynamic single channel data [4] and given in Table 2, whereas Inline graphic is not clearly revealed by experimental data. However we have shown that it should be relatively large for high Inline graphic but relatively small for low Inline graphic for reproducing experimental puff data [8]. By introducing two Inline graphic concentrations, Inline graphic and Inline graphic, Inline graphic and the state of the Inline graphic channel become highly correlated, so that we can assume Inline graphic is a relatively large value Inline graphic if the channel is open and is a relatively small value Inline graphic if the channel is closed. Hence, Inline graphic is modeled by the logic function

graphic file with name pcbi.1003783.e529.jpg

Values of Inline graphic and Inline graphic are chosen so that simulated Inline graphic oscillations in ASMC are comparable to experimental observations.

The Inline graphic model reduction

Here we reduce the 6-state model to a 2-state open/closed model. The reduction takes the following steps:

  • The sum of the probabilities of Inline graphic, Inline graphic and Inline graphic is less than 0.03 for any Inline graphic, so they are either rarely visited by the Inline graphic or have a very short dwell time. This implies they have very little contribution to the Inline graphic dynamics. Therefore, we completely remove the three states from the full model.

  • Transition rates of Inline graphic and Inline graphic are about 2 orders larger than that of Inline graphic and Inline graphic, which allows us to omit the fast transitions by taking a quasi-steady state approximation. This change will affect two aspects. First, we have Inline graphic which allows us to combine Inline graphic and Inline graphic to be a new state Inline graphic, which satisfies Inline graphic. Although this means Inline graphic is a partially open state with an open probability of Inline graphic, it can be used as an fully open state in the stochastic simulations by multiplying the maximum Inline graphic flux conductance Inline graphic by a factor of Inline graphic. Secondly, Inline graphic needs to be rescaled by Inline graphic, i.e., the effective closing rate is Inline graphic.

  • Due to the combination of Inline graphic and Inline graphic, Inline graphic is accordingly modified to

    graphic file with name pcbi.1003783.e560.jpg

Hence, the reduced two-state model contains one “open” state Inline graphic and one closed state Inline graphic with the opening transition rate of Inline graphic and the closing transition rate of Inline graphic.

Deterministic formulation of the stochastic model

Based on the stochastic calcium model and the reduced 2-state Inline graphic model, we construct a deterministic model. We need to modify three things that are used in the stochastic model but inapplicable to fast simulations of the deterministic model. The first is the discrete number of open channels; the second is state-dependent use of Inline graphic and Inline graphic in calculating Inline graphic and Inline graphic; the last is the logic expression of Inline graphic. Details of the modifications are as follows,

  • The fraction of open channels (Inline graphic) is replaced by open probability Inline graphic which is 70% of the probability of state Inline graphic.

  • In the stochastic simulations, Inline graphic which only controls the Inline graphic closing is primarily governed by Inline graphic, whereas Inline graphic which controls Inline graphic opening is mainly governed by Inline graphic. Therefore, in the deterministic model, we separate the functions of Inline graphic and Inline graphic by assuming Inline graphic and Inline graphic are functions of Inline graphic only whereas Inline graphic and Inline graphic are functions of Inline graphic only. That is, Inline graphic, Inline graphic, Inline graphic and Inline graphic. Here Inline graphic as defined before.

  • To describe an average rate that infinitely many receptors are rapidly inhibited by high Inline graphic concentration but slowly restored from Inline graphic-inhibition. Inline graphic is proposed to be

    graphic file with name pcbi.1003783.e596.jpg

Based on the above changes, the full deterministic model containing 8 ODEs is presented as follows,

graphic file with name pcbi.1003783.e597.jpg (13)
graphic file with name pcbi.1003783.e598.jpg (14)
graphic file with name pcbi.1003783.e599.jpg (15)
graphic file with name pcbi.1003783.e600.jpg (16)
graphic file with name pcbi.1003783.e601.jpg (17)
graphic file with name pcbi.1003783.e602.jpg (18)
graphic file with name pcbi.1003783.e603.jpg (19)
graphic file with name pcbi.1003783.e604.jpg (20)

where Inline graphic and Inline graphic are functions of the gating variables given by Eqs. 4 and 5. All the fluxes are the same as those of the stochastic model except Inline graphic. All the parameter values of the deterministic model are the same as those of the stochastic model and are therefore given in Tables 1 and 2.

Reduction of the full deterministic model

The full deterministic model contains 8 variables which make the model difficult to implement and analyze. Thus, we reduce the full model to a minimal model that still captures the crucial features of the full model. First of all, Inline graphic, Inline graphic and Inline graphic are sufficiently large so that we can assume they instantaneously follow their equilibrium functions. Therefore, by taking quasi-steady state approximation to Inline graphic, Inline graphic and Inline graphic, we remove the three time-dependent variables from the full model.

By now, the full model has been reduced to a 5D model,

graphic file with name pcbi.1003783.e614.jpg (21)
graphic file with name pcbi.1003783.e615.jpg (22)
graphic file with name pcbi.1003783.e616.jpg (23)
graphic file with name pcbi.1003783.e617.jpg (24)
graphic file with name pcbi.1003783.e618.jpg (25)

Second, the rate of change of Inline graphic approaching its equilibrium, Inline graphic (calculated from Eq. 24), is at least one order larger than those of Inline graphic, Inline graphic and Inline graphic, indicating that taking the quasi-steady state approximation to Eq. 24 could not significantly affect the evolutions of Inline graphic, Inline graphic and Inline graphic. That is,

graphic file with name pcbi.1003783.e627.jpg (26)

We emphasize here that the theory of the quasi-steady state approximation has not yet been well established, particularly about the rigorous conditions under which such a reduction is valid. Thus, our criterion of judging the validity of the reduction is checking whether the solutions of the reduced model are capable of qualitatively reproducing that of its original model. For this model, we find the reduction works. Hence, the full model is eventually reduced to a 4D model summarized as follows,

graphic file with name pcbi.1003783.e628.jpg (27)
graphic file with name pcbi.1003783.e629.jpg (28)
graphic file with name pcbi.1003783.e630.jpg (29)
graphic file with name pcbi.1003783.e631.jpg (30)

where Inline graphic is given by Eq. 26.

Inclusion of calcium buffers

To check the effect of calcium buffers on oscillation frequency, we introduce a stationary buffer (no buffer diffusion), as mobile buffers are too complicated to be included in the current deterministic model. Since we have two different cytoplasmic Inline graphic concentrations, Inline graphic and Inline graphic, two pools of buffer with the same kinetics should be considered. Hence, the inclusion of a stationary calcium buffer is modeled by the following system,

graphic file with name pcbi.1003783.e636.jpg (31)
graphic file with name pcbi.1003783.e637.jpg (32)
graphic file with name pcbi.1003783.e638.jpg (33)
graphic file with name pcbi.1003783.e639.jpg (34)
graphic file with name pcbi.1003783.e640.jpg (35)

where Inline graphic (Inline graphic and Inline graphic) and Inline graphic represent the concentrations of Inline graphic-bound buffer and total buffer respectively. Inline graphic and Inline graphic are the rates of Inline graphic-binding and Inline graphic-dissociation, indicating how fast the time scale of the buffer dynamics is. Fast buffer refers to the buffer with relatively large Inline graphic. In the simulations, we use a fast buffer with Inline graphic and Inline graphic and vary Inline graphic to test if the stochastic model and the deterministic model have a qualitatively similar Inline graphic-dependency. Results are given in Fig. 6E.

Numerical methods and tools for deterministic and stochastic simulations

For the stochastic model, Eqs. 1–3 and ODEs of the four gating variables in the Inline graphic model are solved by the fourth-order Runge-Kutta method (RK4) and the stochastic states of Inline graphic determined by the Inline graphic model are solved by using a hybrid Gillespie method with adaptive timing [37]. The maximum time step size is set to be either Inline graphic (for the 6-state Inline graphic model) or Inline graphic (for the reduced 2-state Inline graphic model). All the computations are done with MATLAB (The MathWorks, Natick, MA) and the codes are provided in Supporting information (Text S1S2). For the deterministic model, we use ode15s, an ODE solver in MATLAB. Accuracy is controlled by setting an absolute tolerance of Inline graphic applied to all the variables.

Statistical analysis

Data analysis is performed on the Inline graphic traces with relatively stable baselines and less noise. A moving average of every 3 data points is used to improve the data by smoothing out short-term fluctuations (Fig. 2A is an improved result). Due to large variations in baseline, amplitude, and level of noise in data, we used two thresholds to get samples: a low threshold, 20% of the amplitude of the largest spike above the baseline, to initially filter baseline noise out; and a relatively high threshold, 50% of the amplitude of the largest spike above the baseline, to further remove small spikes that cannot initiate waves. For simulated stochastic traces of variable Inline graphic, we first convert it to fluorescence ratio (Inline graphic) by using Inline graphic where the dissociation constant of Oregon Green Inline graphic and resting Inline graphic Inline graphic. We then used the same sampling procedure mentioned above to obtain samples. After samples are chosen, ISIs and spike durations are calculated based on the low threshold. Simulated traces used to calculate average frequency are about 200–400 seconds long. All the samplings and linear least-squares fittings are implemented using MATLAB (see Text S3S4 for Matlab codes).

Supporting Information

Dataset S1

ASMC calcium fluorescence trace data. The data files are in Excel format and compressed in a zip file. Each Excel file has a name showing their information. For example, “S2_SMC6_MCh200nM” means data are from ASMC No. 6 in lung slice No. 2 by using 200 nM MCh. In each file, there are four columns which represent (from left to right) time(s), fluorescence intensity, Inline graphic and average Inline graphic.

(ZIP)

Text S1

Matlab code for simulation using 6 state Inline graphic model.

(DOCX)

Text S2

Matlab code for simulation using 2 state Inline graphic model.

(DOCX)

Text S3

Matlab code for experimental data analysis.

(DOCX)

Text S4

Matlab code for simulation analysis.

(DOCX)

Acknowledgments

We acknowledge many useful conversations with Martin Falcke.

Data Availability

The authors confirm that all data underlying the findings are fully available without restriction. All relevant data are within the paper and its Supporting Information files.

Funding Statement

Funding for all authors came from National Heart Lung Blood Institute (USA) RO1 HL103405. http://www.nhlbi.nih.gov/ The funders had no role in study design, data collection and analysis, decisions to publish, or preparation of the manuscript.

References

  • 1. De Young GW, Keizer J (1992) A single-pool inositol 1, 4, 5-trisphosphate-receptor-based model for agonist-stimulated oscillations in Ca2+ concentration. Proc Natl Acad Sci USA 89: 9895–9899. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2. Dupont G, Goldbeter A (1993) One-pool model for Ca2+ oscillations involving Ca2+ and inositol 1,4,5-trisphosphate as co-agonists for Ca2+ release. Cell Calcium 14: 311–322. [DOI] [PubMed] [Google Scholar]
  • 3. Atri A, Amundson J, Clapham D, Sneyd J (1993) A single-pool model for intracellular calcium oscillations and waves in the Xenopus laevis oocyte. Biophys J 65: 1727–1739. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4. Mak DOD, Pearson JE, Loong KPC, Datta S, Fernández-Mongil M, et al. (2007) Rapid ligand-regulated gating kinetics of single IP3R Ca2+ release channels. EMBO Rep 8: 1044–1051. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5. Wagner LE, Yule DI (2012) Differential regulation of the InsP3 receptor type-1 and -2 single channel properties by InsP3, Ca2+ and atp. J Physiol 590: 3245–3259. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6. Ullah G, Mak DOD, Pearson JE (2012) A data-driven model of a modal gated ion channel: the inositol 1,4,5-trisphosphate receptor in insect sf9 cells. J Gen Physiol 140: 159–173. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7. Siekmann I, Wagner LE, Yule DI, Crampin EJ, Sneyd J (2012) A kinetic model for IP3R type i and type ii accounting for mode changes. Biophys J 103: 658–668. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8. Cao P, Donovan G, Falcke M, Sneyd J (2013) A stochastic model of calcium puffs based on single-channel data. Biophys J 105: 1133–1142. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9. Smith IF, Parker I (2009) Imaging the quantal substructure of single IP3R channel activity during Ca2+ puffs in intact mammalian cells. Proc Natl Acad Sci USA 106: 6404–6409. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10. Brumen M, Fajmut A, Dobovišek A, Roux E (2005) Mathematical modelling of Ca2+ oscillations in airway smooth muscle cells. J Biol Phys 31: 515–524. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11. Sneyd J, Tsaneva-Atanasova K, Reznikov V, Bai Y, Sanderson MJ, et al. (2006) A method for determining the dependence of calcium oscillations on inositol trisphosphate oscillations. Proc Natl Acad Sci USA 103: 1675–1680. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12. Wang IY, Bai Y, Sanderson MJ, Sneyd J (2010) A mathematical analysis of agonist- and kcl-induced Ca2+ oscillations in mouse airway smooth muscle cells. Biophys J 98: 1170–1181. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13. Croisier H, Tan X, Perez-Zoghbi JF, Sandrson MJ, Sneyd J, et al. (2013) Activation of store-operated calcium entry in airway smooth muscle cells: insight from a mathematical model. PLoS ONE 8(7): e69598 10.1371/journal.pone.0069598 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14. Marchant JS, Parker I (2001) Role of elementary Ca2+ puffs in generating repetitive Ca2+ oscillations. EMBO J 20: 65–76. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15. Skupin A, Kettenmann H, Winkler U, Wartenberg M, Sauer H, et al. (2008) How does intracellular Ca2+ oscillate: by chance or by the clock? Biophys J 94: 2404–2411. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16. Perez JF, Sanderson MJ (2005) The frequency of calcium oscillations induced by 5-ht, ach, and kcl determine the contraction of smooth muscle cells of intrapulmonary bronchioles. J Gen Physiol 125: 535–553. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17. Bai Y, Edelmann M, Sanderson MJ (2009) The contribution of inositol 1,4,5-trisphosphate and ryanodine receptors to agonist-induced Ca2+ signaling of airway smooth muscle cells. Am J Physiol Lung Cell Mol Physiol 297: L347–L361. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18. Bird GSJ, Putney JW (2005) Capacitative calcium entry supports calcium oscillations in human embryonic kidney cells. J Physiol 562(3): 697–706. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19. Sneyd J, Tsaneva-Atanasova K, Yule DI, Thompson JL, Shuttleworth TJ (2004) Control of calcium oscillations by membrane fluxes. Proc Natl Acad Sci USA 101: 1392–1396. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20. Mahn K, Hirst SJ, Ying S, Holt MR, Lavender P, et al. (2009) Diminished sarco/endoplasmic reticulum Ca2+ atpase (serca) expression contributes to airway remodelling in bronchial asthma. Proc Natl Acad Sci USA 106: 10775–10780. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21. Sathish V, Leblebici F, Kip SN, Thompson MA, Pabelick CM, et al. (2008) Regulation of sarcoplasmic reticulum Ca2+ reuptake in porcine airway smooth muscle. Am J Physiol Lung Cell Mol Physiol 294: L787–L796. [DOI] [PubMed] [Google Scholar]
  • 22. Zeller S, Rüdiger S, Engel H, Sneyd J, Warnecke G, et al. (2009) Modeling of the modulation by buffers of Ca2+ release through clusters of IP3 receptors. Biophys J 97: 992–1002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23. Li Y, Rinzel J (1994) Equations for InsP3 receptor-mediated [Ca2+]i oscillations derived from a detailed kinetic model: a hodgkin-huxley like formalism. J Theor Biol 166: 461–473. [DOI] [PubMed] [Google Scholar]
  • 24.Keener J, Sneyd J (2009) Mathematical Physiology, Second Edition. Springer, New York.
  • 25. Ionescu L, White C, Cheung KH, Shuai J, Parker I, et al. (2007) Mode switching is the major mechanism of ligand regulation of InsP3 receptor calcium release channels. J Gen Physiol 130: 631–645. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26. Shuai JW, Jung P (2002) Stochastic properties of Ca2+ release of inositol 1,4,5-trisphosphate receptor clusters. Biophys J 83: 87–97. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27. Dupont G, Abou-Lovergne A, Combettes L (2008) Stochastic aspects of oscillatory Ca2+ dynamics in hepatocytes. Biophys J 95: 2193–2202. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28. Kannan MS, Prakash YS, Brenner T, Mickelson JR, Sieck GC (1997) Role of ryanodine receptor channels in Ca2+ oscillations of porcine tracheal smooth muscle. Am J Physiol 272: L659–L664. [DOI] [PubMed] [Google Scholar]
  • 29. Tazzeo T, Zhang Y, Keshavjee S, Janssen LJ (2008) Ryanodine receptors decant internal Ca2+ store in human and bovine airway smooth muscle. Eur Respir J 32: 275–284. [DOI] [PubMed] [Google Scholar]
  • 30. Ressmeyer AR, Bai Y, Delmotte P, Uy KF, Thistlethwate P, et al. (2010) Human airway contraction and formoterol-induced relaxation is determined by Ca2+ oscillations and Ca2+ sensitivity. Am J Respir Cell Mol Biol 43: 179–191. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31. Soeller C, Cannell MB (2004) Analysing cardiac excitation-contraction coupling with mathematical models of local control. Prog Biophys Mol Biol 85: 141–162. [DOI] [PubMed] [Google Scholar]
  • 32. Thurley K, Skupin A, Thul R, Falcke M (2012) Fundamental properties of Ca2+ signals. Biochim Biophys Acta 1820(8): 1185–1194. [DOI] [PubMed] [Google Scholar]
  • 33. Dickinson G, Parker I (2013) Factors determining the recruitment of inositol trisphosphate receptor channels during calcium puffs. Biophys J 105: 2474–2484. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34. Rüdiger S, Shuai JW, Sokolov IM (2010) Law of mass action, detailed balance, and the modeling of calcium puffs. Phys Rev Lett 105(4): 048103 10.1103/PhysRevLett.105.048103 [DOI] [PubMed] [Google Scholar]
  • 35. Rüdiger S, Jung P, Shuai J (2012) Termination of Ca2+ release for clustered IP3R channels. PLoS Comput Biol 8(5): e1002485 10.1371/journal.pcbi.1002485 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36. Chandrasekera PC, Kargacim ME, Deans JP, Lytton J (2009) Determination of apparent calcium affinity for endogenously expressed human sarco(endo)plasmic reticulum calcium-atpase isoform serca3. Am J Physiol Cell Physiol 296: C1105–C1114. [DOI] [PubMed] [Google Scholar]
  • 37. Rüdiger S, Shuai JW, Huisinga W, Nagaiah C, Warnecke G, et al. (2007) Hybrid stochastic and deterministic simulations of calcium blips. Biophys J 93: 1847–1857. [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.

Supplementary Materials

Dataset S1

ASMC calcium fluorescence trace data. The data files are in Excel format and compressed in a zip file. Each Excel file has a name showing their information. For example, “S2_SMC6_MCh200nM” means data are from ASMC No. 6 in lung slice No. 2 by using 200 nM MCh. In each file, there are four columns which represent (from left to right) time(s), fluorescence intensity, Inline graphic and average Inline graphic.

(ZIP)

Text S1

Matlab code for simulation using 6 state Inline graphic model.

(DOCX)

Text S2

Matlab code for simulation using 2 state Inline graphic model.

(DOCX)

Text S3

Matlab code for experimental data analysis.

(DOCX)

Text S4

Matlab code for simulation analysis.

(DOCX)

Data Availability Statement

The authors confirm that all data underlying the findings are fully available without restriction. All relevant data are within the paper and its Supporting Information files.


Articles from PLoS Computational Biology are provided here courtesy of PLOS

RESOURCES