Skip to main content
PLOS One logoLink to PLOS One
. 2013 Apr 5;8(4):e60323. doi: 10.1371/journal.pone.0060323

Regulation of Ion Gradients across Myocardial Ischemic Border Zones: A Biophysical Modelling Analysis

Steven Niederer 1,*
Editor: Vladimir E Bondarenko2
PMCID: PMC3618345  PMID: 23577101

Abstract

The myocardial ischemic border zone is associated with the initiation and sustenance of arrhythmias. The profile of ionic concentrations across the border zone play a significant role in determining cellular electrophysiology and conductivity, yet their spatial-temporal evolution and regulation are not well understood. To investigate the changes in ion concentrations that regulate cellular electrophysiology, a mathematical model of ion movement in the intra and extracellular space in the presence of ionic, potential and material property heterogeneities was developed. The model simulates the spatial and temporal evolution of concentrations of potassium, sodium, chloride, calcium, hydrogen and bicarbonate ions and carbon dioxide across an ischemic border zone. Ischemia was simulated by sodium-potassium pump inhibition, potassium channel activation and respiratory and metabolic acidosis. The model predicted significant disparities in the width of the border zone for each ionic species, with intracellular sodium and extracellular potassium having discordant gradients, facilitating multiple gradients in cellular properties across the border zone. Extracellular potassium was found to have the largest border zone and this was attributed to the voltage dependence of the potassium channels. The model also predicted the efflux of Inline graphic from the ischemic region due to electrogenic drift and diffusion within the intra and extracellular space, respectively, which contributed to Inline graphic depletion in the ischemic region.

Introduction

Myocardial ischemia is caused by reduced perfusion to regions of the heart leading to a localised reduction in supply of metabolites, limited waste removal and compromised ionic homeostasis. The first 10 minutes of ischemia are associated with an increased risk of arrhythmias peaking after 5–6 minutes [1]. During this period arrhythmias are commonly initiated within the border zone (BZ) separating viable, well perfused, tissue and the ischemic, underperfused, region [2][4]. Ischemia causes an increase in extracellular potassium (Inline graphic), intra and extracellular proton concentrations (Inline graphic and Inline graphic, respectively), intracellular sodium (Inline graphic) and intracellular calcium (Inline graphic) concentrations [5]. The dominant mechanisms for these changes have been attributed to a shift in the ATP/ADP ratio, which inhibits the Sodium-Potassium ATPase pump (Inline graphic) and increases the conductance of ATP-inactivated Inline graphic channels; respiratory acidosis causing an increase in Inline graphic; and metabolic acidosis, where a shift towards anaerobic respiration increases the production of Inline graphic in the cell [5]. Inherently, these changes in ionic concentrations in the ischemic region lead to gradients in properties across the BZ, creating electrophysiological heterogeneities that are thought to favour the occurrence of arrhythmias [6][8].

Experimentally, the development of gradients of extracellular pH (Inline graphic) and Inline graphic [9], [10] have been well characterised using ion sensitive electrodes. Intracellular metabolite gradients have been characterised by fluorescent NADH [11] and biopsy [12] measurements. However, less is known on the gradients of intracellular ions, in particular Inline graphic, Inline graphic and Inline graphic, nor are the mechanisms that underpin the spatial and temporal evolution of these ion concentration gradients well characterised or understood. This study aims to investigate the spatial-temporal evolution of ion gradients across ischemic BZ and the primary regulators of the BZ size and rate of development.

Previous measurements of ion concentrations and metabolites across the BZ have either been performed at multiple locations but at a limited number of time points [10], [12] or have tracked the time evolution of ion concentrations but only from a limited number of locations [13], [14]. Furthermore, these measurements have only been able to characterise a subset of ions of interest across the BZ. The need to track the evolution of multiple ionic species in space and time to understand the gradients of cellular electrophysiology across the BZ motivates the use of biophysical computational modelling. Previous models of electrophysiology during acute regional ischemia have simulated the effects of these spatial gradients but have not simulated their time evolution [15][17]. More recent work has simulated the time evolution of Inline graphic gradients [18], but have not considered other ionic gradients, the effects of nonlinear interactions between Inline graphic and other ions, the effect of Inline graphic diffusion in the intracellular space or the effects of potential gradients on ion diffusion.

In this study a new model of cardiac tissue electrophysiology is developed to investigate the spatial-temporal evolution of ionic concentrations across the ischemic BZ, during the first 5 minutes of reduced perfusion. The proposed model extends the conventional bidomain equations to explicitly link membrane potential to ionic concentrations and enforces ionic species conservation. A model of ion regulation across the cell membrane is then developed, parameterized, validated and coupled to the tissue model. This combined model is then used to investigate the spatial and temporal dispersion of ions across the BZ.

Methods

To model the evolution of ionic concentrations in the presence of multiple ionic gradients, electric gradients and heterogeneous tissue properties requires the development of a new set of equations for modelling the myocardium. In the next section the equations to model the BZ are derived and a model of passive cell membrane ion regulation is developed and validated. The changes to the cell membrane model to simulate Inline graphic inhibition, Inline graphic channel activation, respiratory acidosis and metabolic acidosis are then described.

Tissue Model Derivation

Consistent with previous models of cardiac tissue electrophysiology the myocardium is represented as a two phase medium, with each point in the domain containing a fraction of intra and extracellular space. This gives rise to the scaling variables.

graphic file with name pone.0060323.e022.jpg (1)

where Inline graphic is the volume of the Inline graphic space in a unit of Inline graphic myocardium volume, Inline graphic is the volume fraction of space Inline graphic (with Inline graphic corresponding to the extra (Inline graphic) or intra (Inline graphic) cellular space), Inline graphic is the interspace surface area per unit volume myocardium and Inline graphic is area of the interspace surface. The intracellular volume fraction (Inline graphic) can be separated into sub volume fractions representing the cytosol (Inline graphic), mitochondria (Inline graphic) or the sarcoplasmic reticulum (Inline graphic) to model distinct subcellular spaces, as described below. At each point in space there exists an intracellular potential (Inline graphic), an extracellular potential (Inline graphic), a transmembrane potential (Inline graphic) and an intra and extracellular ion concentration for each of the ion species in the model. The movement of each ion species in each domain is driven by diffusion, due to a gradient in ion concentration, and drift, due to a gradient in the electric field. This movement is described by the Nernst-Plank equations

graphic file with name pone.0060323.e040.jpg (2)

where Inline graphic is the concentration of unbound ion Inline graphic in space Inline graphic, Inline graphic is the charge of ion Inline graphic, Inline graphic is the effective diffusion of ion Inline graphic in space Inline graphic, Inline graphic is Faraday's constant, Inline graphic is the gas constant, Inline graphic is the absolute temperature and Inline graphic is the potential in the space Inline graphic. The conventional Nernst-Plank equations are adapted to represent the movement of ions across the cell membrane between the intra and extracellular spaces. This gives

graphic file with name pone.0060323.e054.jpg (3)

where Inline graphic is the flux of ion Inline graphic across the cell membrane. Inline graphic is defined in units of mM msInline graphic mm. The surface separating the two regions is modelled as a simple capacitor and defining the transmembrane potential to scale with intracellular charge gives

graphic file with name pone.0060323.e059.jpg (4)

where Inline graphic is the intracellular charge per unit cell membrane area and Inline graphic is the membrane capacitance per unit cell membrane area. The charge on either side of the membrane is assumed to be equal but opposite, giving

graphic file with name pone.0060323.e062.jpg (5)

Separating the concentration of each ion species in the intra and extracellular space into ions that are membrane bound and make up the membrane charge (Inline graphic) and those that are in solution gives.

graphic file with name pone.0060323.e064.jpg (6)

where Inline graphic and Inline graphic are the ions in solution and membrane bound ions, respectively. By assuming charge neutrality for the ions within the solute gives:

graphic file with name pone.0060323.e067.jpg (7)

Multiplying Eqn 6 by the ion species charge and Faraday's constant, then summing over all ion species in space Inline graphic gives.

graphic file with name pone.0060323.e069.jpg (8)

Converting the concentration of ions per unit volume in domain Inline graphic to charge per unit cell membrane area and introducing a static charge term (Inline graphic) that characterises all charge not attributable to Inline graphic, gives

graphic file with name pone.0060323.e073.jpg (9)

Combining with Eqn 8 gives

graphic file with name pone.0060323.e074.jpg (10)

This defines the charge on the membrane as equal to the unbalanced charge in space Inline graphic. Using the charge balance Eqn 5, gives

graphic file with name pone.0060323.e076.jpg (11)

Differentiating Eqn 11 with respect to time, substituting in Eqn 3 and recognising that all transmembrane fluxes are balanced, provides

graphic file with name pone.0060323.e077.jpg (12)

this ensures that there is no net charge accumulation in any unit volume of myocardium. Defining the relationship between the intra and extracellular potentials gives

graphic file with name pone.0060323.e078.jpg (13)

Using Eqn 4 and the definition of charge (Eqn 10), then gives the algebraic definition of the transmembrane potential:

graphic file with name pone.0060323.e079.jpg (14)

Rearranging Eqn 13 and substituting in the definition of Inline graphic from Eqn 14 allows Inline graphic to be defined in terms of Inline graphic and Inline graphic. Combining this definition of Inline graphic with Eqns 14, 13, 12, 3 and 2 then represents a closed set of equations. These equations are equivalent to the bidomain equations in the case of a single charge carrier and no gradient in ion concentrations, as shown below.

In cardiac myocytes many important ions, including Inline graphic and Inline graphic are heavily buffered, both within the cell and the extracellular space. To account for buffering the free and buffer bound fraction of Inline graphic are calculated. In general, as Inline graphic the effect of ions bound to the cell membrane will not be included in the buffering equations for simplicity. This gives

graphic file with name pone.0060323.e089.jpg (15)

where Inline graphic are the unbound ions and Inline graphic are the ions bound to buffers. At this time all buffers will be treated as rapid and to a single representative buffer species, giving

graphic file with name pone.0060323.e092.jpg (16)

where Inline graphic, Inline graphic and Inline graphic are the concentration, binding affinity and Hill coefficient, respectively, for the buffer of ion Inline graphic in space Inline graphic. Assuming that ions bound to buffers are immobile, only Inline graphic is used to calculate the diffusion and drift of ions in Eqn 3, and similarly in Eqn 12. As binding of ions to a buffer implicitly removes a charged binding site located on a static protein, Eqn 14 remains unchanged.

Due to the complex anatomy of the cardiac myocyte many sub volumes exist within the cell that affect ionic concentrations. The sum of the volume fraction (Inline graphic) values must be less than, but do not have to be equal to, one, allowing the model to represent any volume fractions that are not directly accessible by ions. In particular the Inline graphic variable can represent all space in the cell or can be substituted for Inline graphic representing the volume fraction of the cytosol (a sub volume of Inline graphic). This allows the effects of SR, mitochondrial or other subcellular structure volumes on intracellular ionic concentrations to be accounted for in the model.

Equation summary

The modelled equations are given by

graphic file with name pone.0060323.e103.jpg (17)
graphic file with name pone.0060323.e104.jpg
graphic file with name pone.0060323.e105.jpg
graphic file with name pone.0060323.e106.jpg
graphic file with name pone.0060323.e107.jpg
graphic file with name pone.0060323.e108.jpg

where it is important to note that for non buffered ions Inline graphic and Inline graphic.

Consistency with bidomain equations

Imposing the implicit assumptions of the bidomain equations that charge carriers are not buffered, ion concentrations are homogenous and charge is carried by a single carrier to Eqn 18, the bidomain equations can be derived. Assuming homogenous ion concentrations and considering the case of the intracellular space reduces Eqn 18 to

graphic file with name pone.0060323.e111.jpg (18)

Differentiating Eqn 14 for the intracellular space gives

graphic file with name pone.0060323.e112.jpg (19)

substituting Eqn 18 into Eqn 19 gives

graphic file with name pone.0060323.e113.jpg (20)

As ion concentrations are homogenous conductivity is defined as

graphic file with name pone.0060323.e114.jpg (21)

Then Eqn 20 reduces to

graphic file with name pone.0060323.e115.jpg (22)

Applying the single charge carrier assumption and converting from ionic flux to current gives the first bidomain equation

graphic file with name pone.0060323.e116.jpg (23)

The second bidomain equation is readily derived from applying the homogenous ion concentration assumption to Eqn 12 and multiplying by Faraday's constant (to convert from conserving ion flux to current) giving

graphic file with name pone.0060323.e117.jpg (24)

Substituting in the definition of conductivity from Eqn 21 then gives the second bidomain equation

graphic file with name pone.0060323.e118.jpg (25)

Modelling the Membrane Fluxes

Cardiac electrophysiology is predominantly determined by the movement of Inline graphic, Inline graphic and Inline graphic. For charge neutrality Inline graphic must also be included in the model. To simulate the evolution of acidosis requires the inclusion of Inline graphic, Inline graphic and Inline graphic in the model. All of these ions (and Inline graphic) were modelled in the intra and extracellular space, with Inline graphic and Inline graphic being buffered. The goal of the model, in this study, was not to track the propagation of the action potential but to simulate the gradients of ions that exist over the BZ. These ion gradients were modelled based on the diastolic properties of the cell. This assumption was also a requirement to enable the simulation of minutes, while remaining computationally tractable.

The membrane ion transport pathways are described first for Inline graphic, Inline graphic and Inline graphic. Ion specific channels are then described that balance the flux of each ion species. The channel and transporter densities were determined by imposing zero net flux for each ion species, using the relative densities of Inline graphic and Inline graphic transporters recorded experimentally, and intra and extracellular ionic concentrations and membrane potential values derived from the literature. Where possible experimental data was taken preferentially from rabbit or guinea pig data at body temperature. This limited number of constraints then allowed the model transporters and channel densities to be uniquely determined.

Sodium regulation

The model of Inline graphic regulation included representations of the Inline graphic, sodium calcium exchanger (Inline graphic), sodium hydrogen exchanger (Inline graphic), sodium bicarbonate co-transporter (Inline graphic) and a lumped sodium channel (Inline graphic). Inline graphic was modelled using the thermodynamically consistent equation set proposed by Smith and Crampin [19]. This model was subsequently revised by Terkildsen et al., [20] and this parameter set that was used here. The model for Inline graphic was fitted to guinea pig data as limited rabbit data was available. However, the maximum flux was rescaled to match rabbit data, as described below.

The Inline graphic model was taken from Weber et al., [21]. The model has been fitted to rabbit experimental data at 37°C. The Inline graphic model was based on the model developed by Crampin and Smith [22] and reparameterized by Niederer and Smith [23]. In this study extracellular Inline graphic and Inline graphic regulation of Inline graphic were included. This model was fitted predominantly to sheep Purkinje data [24], although the Inline graphic dependence of Inline graphic remains relatively consistent between species [25]. The Inline graphic model was taken from Crampin and Smith [22]. The model assumes Inline graphic is electro neutral, which is true for only part of the Inline graphic population [25]. There was not sufficient data to fully characterise the electrogenic and electro neutral forms of Inline graphic, hence the electro neutral model was used. Background Inline graphic flux across residual open fast Inline graphic and persistent Inline graphic channels is limited when the cell is quiescent. However, some flux is still present [26] and a simple lumped background ionic flux equation was used, given by

graphic file with name pone.0060323.e156.jpg (26)

to model the residual Inline graphic flux across any open Inline graphic channels. The same equation form was used for modelling all background ion channels.

Proton regulation

In the proposed model Inline graphic was regulated by Inline graphic, described above, chloride-hydroxide exchanger (Inline graphic), hydrolysis and buffering. Background Inline graphic leak or other Inline graphic exchangers were not considered in the general model of Inline graphic regulation, described here, but do include models of Inline graphic-lactate exchange and intracellular metabolism derived Inline graphic sources in the model of ischemia, described below. The Inline graphic model comes from Niederer et al., [27] and was fitted to guinea pig data at 37Inline graphicC. The hydrolysis of Inline graphic into Inline graphic and Inline graphic was governed by

graphic file with name pone.0060323.e172.jpg (27)

where Inline graphic and Inline graphic are the forward and reverse rates of hydrolysis. Hydrolysis occurs in both the intra and extracellular space and the rate constants were assumed to be the same in both domains. Inline graphic buffering in the intra and extracellular space is due to mobile and static Inline graphic buffers and Inline graphic [28]. The intra and extracellular buffering of Inline graphic were assumed to be instantaneous and represented by a single population of buffers. To reduce the model size, partial differential equations were only solved for the total concentration of Inline graphic or Inline graphic. The concentration of free ions were then calculated by

graphic file with name pone.0060323.e181.jpg (28)

where Inline graphic and Inline graphic represent the concentration of the buffer and the binding affinity, respectively. This buffer model was used for both intra and extracellular Inline graphic, with a separate set of parameters for each ion.

Calcium regulation

A simplified model of intracellular cardiac Inline graphic was developed assuming that Inline graphic in the intracellular space reaches an approximate equilibrium over the time scales of interest. Furthermore, SERCA ATPase function was modelled with a Hill coefficient of one as opposed to two, to allow the definition of Inline graphic to remain deterministic. The intracellular space was assumed to consist of a sarcoplasmic reticulum (SR) and a cytosolic space. The subsarcolemmal and dyadic space are small and are also likely to be in equilibrium with the cytosolic Inline graphic, so were not included in the model. The Inline graphic dynamics were described by

graphic file with name pone.0060323.e190.jpg (29)
graphic file with name pone.0060323.e191.jpg (30)
graphic file with name pone.0060323.e192.jpg (31)
graphic file with name pone.0060323.e193.jpg (32)
graphic file with name pone.0060323.e194.jpg (33)

where Inline graphic is the SR Inline graphic, Inline graphic is the flux of calcium out of the SR, Inline graphic is the uptake of Inline graphic by SERCA, Inline graphic is the maximum SERCA flux, Inline graphic is the diffusion permeability of the SR membrane, Inline graphic is the binding coefficient of Inline graphic to SERCA, Inline graphic and Inline graphic are the volume fractions of the SR and cytosol, respectively, Inline graphic is the L-type calcium channel and Inline graphic is the background Inline graphic channel. In this model the background and L-type Inline graphic channels were modelled as a single lumped generic Inline graphic channel. Introducing cytosolic buffering, ignoring the effects of SR buffering and setting Inline graphic as one and assuming that the cytosol and the SR are in equilibrium then gives

graphic file with name pone.0060323.e212.jpg (34)
graphic file with name pone.0060323.e213.jpg

defining

graphic file with name pone.0060323.e214.jpg (35)
graphic file with name pone.0060323.e215.jpg (36)

and collecting terms gives

graphic file with name pone.0060323.e216.jpg (37)
graphic file with name pone.0060323.e217.jpg (38)
graphic file with name pone.0060323.e218.jpg (39)
graphic file with name pone.0060323.e219.jpg (40)

As Inline graphic is always negative the cubic always has at least one positive real root for possible values of Inline graphic. The value of Inline graphic was then found using the root finding method first proposed by Francois Viete in 1600 and reused more recently by Faber and Rudy [29]:

graphic file with name pone.0060323.e223.jpg (41)

The model of Inline graphic dynamics assumes that all Inline graphic buffers were static and that the transport of ions via mobile buffers was accounted for in the effective diffusion parameters of free Inline graphic. It is possible to extend the proposed model to include mobile buffers but they were assumed to play a secondary role in the current model. Inline graphic was assumed to be buffered by a single species and was modelled using the same framework described above for Inline graphic (Eqn 28).

Chloride regulation

In this model Inline graphic homeostasis was maintained by the Inline graphic and the Inline graphic-Inline graphic exchanger (Inline graphic), which bring Inline graphic into the cell, and a Inline graphic channel that allows Inline graphic to flow out of the cell. The Inline graphic model is described above. The Inline graphic model was taken from Crampin and Smith [22] and was developed using guinea pig data at 37Inline graphicC. The Inline graphic channel uses the conventional background channel formulation. A linear Inline graphic dependence of the background Inline graphic current was added to the model based on observations from Komukai et al., [30].

Potassium regulation

In dynamic action potential models of cardiac electrophysiology there are a large number of Inline graphic channels [31] that bring Inline graphic into the cell. This influx was balanced by the Inline graphic pump, described above, that extrudes Inline graphic. For the passive membrane model, all of the Inline graphic channels were lumped into a single background current (Inline graphic) formulation that was set to balance the flux of Inline graphic on Inline graphic. It was assumed that the membrane potential and Inline graphic reversal potential are the dominant factors affecting this channel and other forms of regulation have not been considered.

Bicarbonate regulation

Inline graphic was assumed to be regulated principally by hydrolysis and through Inline graphic and Inline graphic. All of these components have been described above and it was assumed that there are no other Inline graphic pathways across the membrane.

Carbon Dioxide

Inline graphic is regulated primarily through hydrolysis and can diffuse relatively freely across the membrane. The model of hydrolysis is described above and Inline graphic diffusion was assumed to obey Fick's law.

Model Parameters

For each transmembrane ion pathway described above, all kinetic, binding affinity and membrane potential dependencies were taken from the original models. Here the definition of geometrical parameters, ionic concentrations, buffering parameters and the density/scaling of each transmembrane pathway are motivated from data in the literature.

Geometrical parameters

The extracellular space is estimated to be between Inline graphic [32][35], Inline graphic and Inline graphic [33], [36] of the volume of the heart in rabbit, cat, and rat hearts, respectively. This gave an Inline graphic value of Inline graphic leading to an Inline graphic value of Inline graphic. The surface to volume ratio of a cell is reported as Inline graphic mInline graphic [37], [38] in rat and rabbit myocytes, corresponding to a Inline graphic value of Inline graphicmmInline graphic. The relative SR volume was set to Inline graphic of intracellular volume, giving an Inline graphic value of Inline graphic, based on reported values of Inline graphic of cell volume [37], [39], [40] in mouse, rat and swine. The relative mitochondrial volume (Inline graphic) was set to Inline graphic based on an estimated mitochondrial cell volume fraction of Inline graphic [37], [40]. The cytosol volume fraction was set to Inline graphic of the intracellular space, resulting in an Inline graphic value of Inline graphic. To account for the effects of subscellular domains on intracellular ionic concentrations in the model, all references to Inline graphic in Eqn 18 were replaced by Inline graphic. A summary of geometrical parameters is given in Table 1.

Table 1. Geometric variables.
Variable Value
Inline graphic 0.8
Inline graphic 0.2
Inline graphic 264 mmInline graphic
Inline graphic 0.536
Inline graphic 0.24
Inline graphic 0.024

Intracellular ionic concentrations

Inline graphic has been measured using SBF1 fluorescence and Inline graphic sensitive electrodes. A significant range of values have been reported from Inline graphicmM [26], [41][44] to Inline graphicmM [34], [45], [46]. Early measurements of intracellular ionic concentrations were performed using ion sensitive electrodes. These experiments measure ion activity and not ionic concentration and are often performed in multi-cellular preparations, confounding measurements. For these reasons Inline graphic in quiescent myocytes was set to Inline graphicmM, consistent with recent calibrated fluorescent measurements in isolated rabbit myocytes [26], [44].

No fluorescence dye is routinely used for measuring Inline graphic. Using ion sensitive electrodes Lee et al., [45] were able to calibrate their measurements of ion activity in rabbit myocytes using an estimated Inline graphic activity coefficient of Inline graphic, giving a value of Inline graphicmM. This compares with a range of Inline graphicmM calculated by applying the Lee et al., Inline graphic ion activity coefficient to ion activity measurements in rabbit, cat and guinea pig [34], [43], [47], [48]. Alternate measurement using flame emission spectrometry by Powell et al., [49] measured Inline graphic in rat myocytes, giving a concentration of Inline graphic mM. Given the lower values of the two calibrated measurements, Inline graphic was set to Inline graphic mM.

No dye is routinely used for measuring Inline graphic concentration in cardiac cells, however, Inline graphic can be measured using ion sensitive electrodes. Inline graphic activity has been reported as Inline graphic mM [35], [48], [50], [51] in sheep, rabbit and guinea pig heart cells. Estimations of Inline graphic from total tissue Inline graphic concentrations have resulted in values of Inline graphic mM [35] and Inline graphic mM [34] in rabbit cells and Inline graphic mol/g dry wt (or Inline graphic mM using the Inline graphic mM per Inline graphicmol/g dry wt scaling factor from Bers [52]) in rat. The higher value of Inline graphic mM may be attributed to the higher extracellular space used in these calculations (Inline graphic as compared to Inline graphic). Considering the relative convergence of values Inline graphic was set to Inline graphic mM.

Resting free Inline graphic is measured using calibrated fluorescence measurements. These measurements range from Inline graphic nM in rabbit and guinea pig preparations [53][57]. Given this consistency the Inline graphic will be set to Inline graphic nM. SR Inline graphic concentration is calculated from integrating the current across the cell membrane following the release of Inline graphic from the SR in response to caffeine. These measurements show two populations with high values in rat (Inline graphicM [58][61]), canine (Inline graphicM [62]), rabbit (87–106 Inline graphicM [60], [61]) and ferret (Inline graphicM [63]), compared to lower measurements in guinea pig (Inline graphicM [58], [64]). Given that the majority of species have a higher reported concentration, including rabbit, simulations were run with SR Inline graphic load set to Inline graphicM. The buffering of Inline graphic can be described by Hill equation(s), mass action equation(s) or a constant buffering power. To compromise between biophysics and complexity, buffering was modelled by a single Hill equation. Hove-Madsen and Bers [65] fitted Inline graphic buffering in rabbit myocytes using two Hill curves; however, the lower affinity buffer will not play a significant role at passive diastolic Inline graphic concentrations. For this reason the high affinity site, with cooperativity reduced from Inline graphic to unity, was used to model Inline graphic buffering, giving a buffer concentration of 208.98 Inline graphicM (converted using a scaling factor of Inline graphic from Bers [52]) and an affinity of Inline graphicM. This model of Inline graphic buffering in rabbit is similar to the concentration/affinity values of Inline graphicM [66], Inline graphicM [67] and Inline graphicM [68] values measured in other species.

Inline graphic has been measured using Inline graphic sensitive electrodes and fluorescence dyes. Measurement of Inline graphic consistently falls within the range of Inline graphic [57], [69][71] in either HEPES or Inline graphic buffered solutions. In the model Inline graphic was set to 7.1. Inline graphic are heavily buffered in the cytosol by intrinsic buffers and Inline graphic. Here the buffering of Inline graphic by Inline graphic was modelled explicitly and the intrinsic buffers were assumed to be in rapid equilibrium. Leem et al., [70] measured (and modelled) Inline graphic buffering by two populations of buffers with binding affinity pK values of Inline graphic and Inline graphic and concentrations of Inline graphicmM and Inline graphicmM. Zanbioni et al., [25] differentiated between mobile and fixed buffers and found that the fixed buffers had a consistent concentration of Inline graphicmM and binding affinity (pK) value of Inline graphic across rat, rabbit and guinea pig, while the mobile buffers had a constant Inline graphic value of Inline graphic. Fitting a single buffering curve to these two models over a pH range of Inline graphic gives concentrations of Inline graphicmM and pK values of Inline graphic, which gave a value of Inline graphicmM and pK value of Inline graphic for this model.

Extracellular ionic concentrations

The concentration of the majority of ions in the extracellular space have been measured in canine [46], [72], rat [36], [73] and cat [74] hearts. These measurements provide a consistent range of ion concentrations for Inline graphic, Inline graphic and Inline graphic, giving Inline graphic as Inline graphicmM, Inline graphic as Inline graphicmM and Inline graphic as Inline graphicmM. In the model Inline graphic was set to Inline graphicmM, Inline graphic was set to Inline graphicmM, consistent with measurements in guinea pig hearts [75] and rabbit atrium [76] and Inline graphic was set to Inline graphicmM. The Inline graphic values reported range from Inline graphicmM, however, these values do not differentiate between buffered and ionized Inline graphic. The properties of Inline graphic buffering in the extracellular space are not well characterised and are generally ignored in previous cardiac cell models. To approximate the buffering properties of extracellular Inline graphic using a simple single species steady state mass action model, with no cooperative binding, requires two parameters, the concentration of the buffer and the binding affinity. Assuming the ratio of free Inline graphic in the extracellular space to bound ions is similar to serum [77] and assuming that the primary buffer of Inline graphic in the extracellular space are phospholipids, then extracellular Inline graphic buffering will have a binding affinity of Inline graphicmM [78], within the range observed across multiple species [79], and a buffer concentration of Inline graphicmM based on a free Inline graphic concentration of Inline graphicmM and assuming Inline graphic of Inline graphic are bound to buffers [77]. Inline graphic was set to Inline graphic, to be consistent with the majority of experimental studies [57], [69][71] and measurements across a range of species [80]. Limited measurements were available to model Inline graphic buffering, however, Yan and Kleber [81] reported 39 mM of buffered Inline graphic in the extracellular space. By assuming similar binding affinities for the intra and extracellular buffers and that Inline graphic is Inline graphic, then gave a concentration of extracellular proton buffers of Inline graphicmM. A summary of ion concentrations and buffering parameters are given in Table 2 and Table 3, respectively.

Table 2. Intra and extracellular free ion concentrations.
Ion Concentration (mM)
Intracellular Extracellular
Inline graphic 4.0 140
Inline graphic 135 4.0
Inline graphic 18 110
Inline graphic Inline graphic 1.2
Inline graphic Inline graphic Inline graphic
Inline graphic 1.17 1.17
Table 3. Buffering Parameters.
Parameter Value (mM)
Intracellular Extracellular
Inline graphic 65 350
Inline graphic Inline graphic Inline graphic
Inline graphic 0.209 2.3
Inline graphic Inline graphic 1.1

The concentration of Inline graphic was calculated using the parameters proposed and measured for guinea pig ventricular myocytes at Inline graphicC by Leem and Vaughan-Jones [82]. The concentration of Inline graphic in the extracellular solution was calculated using

graphic file with name pone.0060323.e426.jpg (42)

where Inline graphic is the solubility of Inline graphic, set to Inline graphic mM mmHgInline graphic from human measurements at pH Inline graphic and Inline graphicC [83], Inline graphic is the fraction of air that is Inline graphic, set to Inline graphic at baseline and Inline graphic is atmospheric pressure, set to Inline graphic mmHgInline graphic. This gives a partial pressure of Inline graphic (Inline graphic) of Inline graphic mmHg, consistent with although slightly higher than the Inline graphic3 mmHg measured in rabbit hearts [84]. The hydration of Inline graphic was modelled by a mass action reaction (Eqn 27). The Inline graphic and Inline graphic values were measured and modelled by Leem and Vaughan-Jones in guinea pig myocytes at Inline graphicC [82], giving values of Inline graphicsInline graphic and Inline graphicsInline graphic, respectively, resulting in an equilibrium constant of Inline graphic. The rate of hydration of Inline graphic was assumed to be similar in both the intra and extracellular space. The movement of Inline graphic across the cell membrane was modelled by Fick's law. Previous models have used permittivity values of Inline graphicmmsInline graphic [82], based on measurements in red blood cells. This value was reused despite its lack of species and cell type consistency, as there were no recent studies characterising the permeability in cardiac myocytes and the high permeability leads Inline graphic to be close to equilibrium between the intra and extracellular spaces.

Current, exchanger and pump densities

The model of each transmembrane ion pathway (Inline graphic) was separated into a kinetic regulatory component (Inline graphic) dependent on transmembrane potential and ionic concentrations, and a scalar (Inline graphic) representing either the maximum flux or channel conduction of the pathway. The flux across a pathway (Inline graphic) is then given by

graphic file with name pone.0060323.e461.jpg (43)

The Inline graphic values were determind by calculating all of the Inline graphic values (excluding Inline graphic, Inline graphic and Inline graphic as Inline graphic and Inline graphic were unknown) using the ionic concentrations in Table 2 and assuming a membrane potential of −80 mV. The remaining unknown Inline graphic and ionic concentrations were then determined from a limited number of measurements and by enforcing a zero net flux condition described by

graphic file with name pone.0060323.e470.jpg (44)

As the parameters for hydrolysis and permeability of the cell membrane to Inline graphic (Inline graphic, Inline graphic and Inline graphic) were derived from the literature, the constraints on Inline graphic and Inline graphic were used to determine the concentrations of Inline graphic and Inline graphic, respectively. Measurements of transmembrane Inline graphic influx gave Inline graphic as Inline graphic mM msInline graphic, Inline graphic as Inline graphic mM msInline graphic, Inline graphic as Inline graphic mM msInline graphic and Inline graphic as Inline graphic mM msInline graphic [85]. The Inline graphic Inline graphic flux was assumed to be equal to the sum of all Inline graphic influx. As the model does not include Inline graphic-Inline graphic-2-Inline graphic co-transporter (NaK2Cl) or Inline graphic-Inline graphic exchanger (NaMg), primarily due to the limited data to constrain the kinetics, Inline graphic was reduced from the flux measured by Despa et al., [85] to 1.13Inline graphic mM msInline graphic. Setting these Inline graphic values defines Inline graphic, Inline graphic, Inline graphic, Inline graphic, Inline graphic and Inline graphic. In the experimental and modelling work by the Vaughan-Jones group [70], [86] the relative size of Inline graphic to Inline graphic is Inline graphic at Inline graphic, this gave an estimated ratio of Inline graphic. Scaling Inline graphic determined from [85] by 0.16 provided an estimate of Inline graphic of Inline graphic mM msInline graphic. This allows Inline graphic and Inline graphic to be defined. Inline graphic must be balanced by Inline graphic and as Inline graphic and Inline graphic were known, this flux was used to set Inline graphic. Combining the Inline graphic and Inline graphic concentrations with the defined hydrolysis parameters and the known value of Inline graphic then gave Inline graphic. Knowing Inline graphic and Inline graphic gave Inline graphic and hence Inline graphic. Similarly, Inline graphic was calculated from Inline graphic and Inline graphic,which gave Inline graphic. Finally, Inline graphic was calculated using Inline graphic and Inline graphic, calculated using the derived Inline graphic value. By automating this parameter derivation process the model parameters could be updated to ensure static ion concentrations, and hence membrane potential, for any perturbation in model parameters or ionic concentrations. Parameters for all simulations in this study were derived following this process.

Intracellular calcium dynamics

The intracellular regulation of Inline graphic was treated as an equilibrium system. This resulted in the SR effectively acting as an additional buffer on Inline graphic. The parameters Inline graphic, Inline graphic, Inline graphic, Inline graphic and Inline graphic were defined from enforcing a zero net flux constraint and experimental measurements. Measurements of Inline graphic are similar in rat and rabbit [65] and have been reported as 0.260–0.350Inline graphicM in rabbit [65], [87] and 250–280Inline graphicM in rat [65], [87], [88]. A value of 0.3Inline graphicM was used in the model. Inline graphic is reported as 0.001–0.012 Inline graphicMmsInline graphic in rabbit, mouse and rat [54], [89][91]. A value of 0.01Inline graphicMmsInline graphic was used in the model. Enforcing a zero net flux constraint on the SR gave Inline graphic as 0.038 Inline graphicMmsInline graphic, comparable with values of 0.03–0.08Inline graphicMmsInline graphic measured in rabbit [65], [87], [92] with units converted using scale factors from Bers [52].

Membrane Model Validation

To validate the passive membrane model, a series of tests based on the response of the cell models intracellular ionic concentrations and transmembrane potential to changes in extracellular ionic concentrations or inhibition of major ion transporters was performed. In order to maximise the number of tests the model was compared against data from cardiac cells, regardless of temperature, species or preparation type. This maximised the number of tests but meant that a quantitative comparison was not valid and so only a qualitative comparison was performed. A summary of experimental data used in the validation is provided in Table 4. Fig. 1 shows the comparison between the model and data. From the 72 simulations performed 43 matched the experimental data, data was not available or was inconsistent for 19, changes were too small to be measured in 5 and the model did not match experiments in 5 cases.

Table 4. Qualitative changes in intracellular ionic concentrations or membrane potential in response to changes in extracellular ionic concentration or inhibition of membrane transporters.

Protocol Change in Concentration or Potential Reference
Variable Change Inline graphic Inline graphic Inline graphic Inline graphic Inline graphic Inline graphic
1-24-9 Inline graphic
Decrease Inline graphic Inline graphic Inline graphic Inline graphic Inline graphicx2 Inline graphicx4 [160][163]
[164][166]
Increase Inline graphic Inline graphic Inline graphic Inline graphic [167], [168]
Inline graphic
Decrease Inline graphic Inline graphic Inline graphic Inline graphic Inline graphic Inline graphic [163], [169], [170]
Increase Inline graphic Inline graphic Inline graphic [171], [172]
Inline graphic
Decrease Inline graphic Inline graphic Inline graphic Inline graphic Inline graphic Inline graphic Inline graphic [160], [161], [173], [174]
[71], [162], [175], [176]
Increase Inline graphic Inline graphic Inline graphic Inline graphic Inline graphic Inline graphic Inline graphic Inline graphic Inline graphic [50], [161], [175], [177], [178]
[131], [173], [176], [179], [180]
Inline graphic
Increase Inline graphic Inline graphic Inline graphic Inline graphic [71], [166], [181][183]
Inline graphic
Decrease Inline graphic Inline graphic Inline graphic Inline graphic Inline graphic [184][186]
[50], [71], [187]
Increase Inline graphic Inline graphic Inline graphic Inline graphic [50], [184], [187]
Inline graphic
Increase Inline graphic Inline graphic Inline graphic Inline graphic Inline graphic Inline graphic [161], [165], [166], [188]
Inline graphic
Inhibition Inline graphic Inline graphic Inline graphic Inline graphic Inline graphic Inline graphic Inline graphic [43], [170], [175]
[35], [160], [177], [189]
Inline graphic
Inhibition Inline graphic Inline graphic Inline graphic Inline graphic [44], [166]

Figure 1. Cell membrane model validation.

Figure 1

Arrow direction shows model prediction for the change in intracellular concentrations and membrane potential for a change in extracellular ion concentrations or inhibition of membrane transporter as indicated by the panel label. The color of the arrow indicates how the model compares with experimental data, summarised in Table 4. Black arrows indicate where the model matches experimental data, white arrows indicate where there is no or inconsistent data, gray arrows indicate where no change was observed experimentally and striped arrows indicate where the model does not match experimental data. Data was considered inconsistent if different studies reported opposite changes in intracellular ion concentration or membrane potential in response to a change in extracellular ionic concentrations (for example the change in membrane potential following a decrease in Inline graphic). In cases where one study reported no change in intracellular ion concentration or membrane potential and another found a change, it was assumed that the change was correct (for example the change in Inline graphic in response to an increase or a decrease in Inline graphic). For example in panel A) corresponding to a decrease in Inline graphic (as indicated by the panel label), the experimental data comes from row 1 of Table? 4. The model predicts that Inline graphic decreases and that Inline graphic and Inline graphic increase, consistent with experimental measurements and the arrows are shaded black. There is no consistent observed change in the transmembrane potential so the Inline graphic arrow remains white. No data is available for the change in Inline graphic in response to a decrease in Inline graphic so the Inline graphic arrow remains white. The model predicts a decrease in Inline graphic yet experimental measurements found an increase, hence the Inline graphic arrow is striped.

Diffusion Parameters

The tissue model required the definition of diffusion parameters for each of the ionic species in the intra and extracellular space. The diffusion parameters used in the model were all for the apparent diffusivity of free ions, lumping the effects of any mobile buffers, tortuosity and gap junctions into a single diffusion parameter. Using Eqn 21 the conductivity parameters were shown to be directly related to the conductivity parameters in the bidomain equations. From the review of bidomain conductivities and relative values in the intra and extracellular space by Roth [93] a conductivity value of 0.25 SmInline graphic was taken for both the intra and extracellular space. This conductivity, along with the ionic concentrations from Table 2, were used to derive the diffusion parameters. The apparent diffusion of Inline graphic and Inline graphic are Inline graphicmmInline graphicmsInline graphic and Inline graphicmmInline graphicmsInline graphic in the >intracellular space [86] and were assumed to be similar in the extracellular space. The apparent diffusion constant of Inline graphic (including the effects of mobile buffers) has been estimated from experimental measurements in previous modelling studies as Inline graphicmmInline graphicmsInline graphic [94][96]. Here a value of Inline graphicmmInline graphicmsInline graphic [94] was used and the effect of diffusion in the SR was not included, as it was assumed to be non contiguous between cells. It was assumed that Inline graphic diffusion is limited by buffering, resulting in the same value in the intra and extracellular space. Inline graphic, Inline graphic and Inline graphic were all assumed to have a common diffusion coefficient in either the intra or extracellular space as these ions are only nominally buffered and are only affected by gap junctions and tortuosity. Diffusion of Inline graphic was estimated from previous modelling/experimental studies of Inline graphic in tissue as Inline graphicmmInline graphicmsInline graphic [97], and was assumed to be the same in the intra and extracellular space. Solving Eqn 21 for the diffusion of intra and extracellular Inline graphic, Inline graphic and Inline graphic gives the diffusion parameters summarised in Table 5.

Table 5. Effective ion diffusion parameters.

Ion Diffusion Constant (10Inline graphicmmInline graphicmsInline graphic)
Intracellular Extracellular
Inline graphic 7.7 12.9
Inline graphic 7.7 12.9
Inline graphic 7.7 12.9
Inline graphic 1.5 1.5
Inline graphic 1.52 1.52
Inline graphic 7.7 7.70
Inline graphic 11.3 11.3

Finally, the effects of Inline graphic on gap junction conductivity were included. Measurements in cell pairs by Swietach et al., [98] have demonstrated that the permeability of gap junctions has a biphasic dependence on the Inline graphic concentration. To introduce these effects into the model the intracellular diffusion constants were scaled by

graphic file with name pone.0060323.e695.jpg (45)

where Inline graphic/Inline graphic and Inline graphic/Inline graphic are the cooperativity and binding affinity for the activation and deactivation of gap junctions by protons, respectively and Inline graphic scales all intracellular diffusion constants. Using the measured parameters from end-to-end cell pairs in Swietach et al., [98] Inline graphic, Inline graphic, Inline graphicmM and Inline graphicmM. In this model of proton effects on gap junction permeability there is an implicit assumption that the gap junction permeability plays a dominant role in defining diffusion. At this stage no other regulators of permeability were included in the model, notably Inline graphic regulation is absent but this can readily be included in the modelling framework as required.

Simulating Ischemia

In this study the effect of Inline graphic inhibition, Inline graphic activation and respiratory and metabolic acidosis during ischemia on cell ionic homeostasis were considered. This list is not exhaustive and absent factors are discussed below. Ischemia was modelled by respiratory acidosis, metabolic acidosis, increased Inline graphic channel conductance and decreased Inline graphic function. The specific time course of each of these changes is poorly characterised. Previous modelling studies have assumed that changes in cell function with ischemia have evolved linearly [99] or as a nonlinear function of prescribed metabolite concentrations [20]. To avoid any undue bias from the arbitrary selection of a time course, all changes are initially considered instantaneous.

NaK inhibition and Inline graphic channel activation were modelled by scaling the respective fluxes. Respiratory acidosis was assumed to result from an imbalance of production and washout of Inline graphic. To simulate this, an (implicitly electro neutral) intracellular source of Inline graphic was introduced into the ischemic region. There was no flux of Inline graphic out of the extracellular space and this resulted in a build up of Inline graphic in both the intra and extracellular space in the ischemic region.

To model metabolic acidosis required the introduction of an additional intracellular Inline graphic source. However, introducing a source of cations into the cell compromised the conservation of charge constraint. To provide an electro neutral source of Inline graphic required the concurrent introduction of a source of anions that match the production of Inline graphic inside the cell. It was implicitly assumed that the anions entered the cell in an electro neutral form with an Inline graphic bound, and the anions and Inline graphic separate within the cell due to metabolic processes. The source of Inline graphic in ischemia is likely to be due to increased ATP production through glycolysis [100]; in the absence of oxygen this also results in increased lactate production. As lactate readily dissociates from Inline graphic it was assumed that the anion source that matches Inline graphic flux has similar characteristics to lactate including transmembrane regulation via the lactate-Inline graphic membrane exchanger MCT1 [101]. Given the ambiguity in structure, MCT1 was modelled as an ordered exchanger, using the kinetic parameters from Vinnakota and Beard [102] and the maximum influx value of Inline graphicmMmsInline graphic recorded in guinea pig myocytes [103]. In the absence of metabolic acidosis, lactate concentration was assumed to be nominal, consistent with the small flux of lactate observed in rabbit hearts under normal conditions [104]. Model simulations were performed on a 1D strand with length 32 mm, oriented in the preferred conduction or fibre direction, with a transition between ischemic and viable tissue at 16 mm to ensure the simulation captured the BZ width with nominal boundary condition artefacts. The transition between viable tissue and ischemic tissue was approximated by a Hill equation with a Hill coefficient set to ensure a steep transition between the ischemic region and viable tissue over Inline graphicm, consistent with the rapid drop in oxygen pressure across the BZ in swine [105].

Numerical Methods for Tissue Model

The nonlinear equations were solved using a fully implicit finite difference scheme with a line search Newton-Raphson method. The transmembrane flux component of the Jacobian was calculated using finite differencing with the remainder calculated analytically. The Jacobian was inverted directly using MatLab and was only recalculated if the residual fails to decrease or convergence was not reached within Inline graphic Newton-Raphson iterations.

The dependence of the BZ width on the spatial and temporal discretizations was determined to test for numerical convergence. The width of the transition for each ion concentration between the viable and ischemic region, referred to as the BZ width for each ion, was calculated by fitting a Hill curve to each ionic profile. The width of the BZ for each ion was then calculated as the distance between Inline graphic and Inline graphic of the change in concentration.

A convergence analysis was performed and a mesh discretization of Inline graphicm and a time step of Inline graphicms was used. Increasing the spatial discretization by a factor of Inline graphic results in a maximum change in BZ width of any ion of Inline graphicmm. Decreasing the time step to Inline graphicms increases the maximum BZ width of any ion by Inline graphicmm.

Results

The individual effects of each of the four components of ischemia were first demonstrated. A combined model of ischemia was then developed and the width and magnitude of the changes in ionic concentrations across the ischemic BZ were predicted. The effect of movement of Inline graphic, Inline graphic and Inline graphic within and between intra and extracellular spaces across the BZ were calculated to show the net movement of ions across the BZ.

Simulating Individual Components of Ischemia

The level of inhibition of Inline graphic, the activation of Inline graphic and the level of Inline graphic and Inline graphic production in the cell during ischemia is not known. To investigate the effects of each of these aspects of ischemia, they were each individually introduced into the model at five levels of severity, for xInline graphicmm along a Inline graphicmm strand. Fig. 2 shows the effect of each component on the change in Inline graphic, Inline graphic, Inline graphic and Inline graphic across the ischemic BZ. The range of alterations in ion concentrations is shown by the shaded regions. The minimum and maximum changes in each component are shown by dashed lines and the mid change is shown by the solid line. Inline graphic (yellow) was inhibited by up to 100% (no flux) in Inline graphic increments, Inline graphic (red) was scaled by up to Inline graphic in Inline graphic increments, metabolic acidosis was simulated by introducing an Inline graphic flux in five increments up to a maximum value of Inline graphicMmsInline graphic and respiratory acidosis (blue) was simulated by a intracellular Inline graphic flux increased in five increments up to a maximum value of Inline graphicMmsInline graphic.

Figure 2. Effect of individual components of ischemia on the change in Inline graphic (column 1), Inline graphic (column 2), Inline graphic (column 3) and Inline graphic (column 4).

Figure 2

Effect of 50–250% increase in Inline graphic (red shaded region enclosed by dashed line) and a 150% increase in Inline graphic (red line) on A) Inline graphic, B) Inline graphic, C) Inline graphic and D) Inline graphic. Effect of 20–100% maximum Inline graphic flux (purple shaded region enclosed by dashed line) and 60% maximum Inline graphic flux (purple line) on E) Inline graphic, F) Inline graphic, G) Inline graphic and H) Inline graphic. Effect of 20–100% NaK inhibition (yellow shaded region enclosed by dashed line) and a 60% NaK inhibition (yellow line) on I) Inline graphic, J) Inline graphic, K) Inline graphic and L) Inline graphic. Effect of 20–100% maximum Inline graphic flux (blue shaded region enclosed by dashed line) and a 60% maximum Inline graphic flux (blue line) on M) Inline graphic, N) Inline graphic, O) Inline graphic and P) Inline graphic.

Modelling Ischemia

Inherently, there is no single mode of ischemia and the relative contribution of acidosis, Inline graphic inhibition or Inline graphic activation will depend on the residual flow, age, gender, disease state, location of ischemic region and species under study. To provide a representative case to study, 5 minutes of ischemia were simulated in the presence of all four ischemic mechanisms that match representative results from the literature.

Partial pressure measurements of Inline graphic in canine ischemic models show a Inline graphic increase in Inline graphic partial pressure following occlusion, depending on the level of flow inhibition, after Inline graphic minutes. The elevation in Inline graphic was approximately linear over time and hence in the model it was assumed Inline graphic concentration increases by Inline graphic increase in the first Inline graphic minutes of ischemia [106], [107]. This corresponded to an increase in the concentration of Inline graphic in the model from Inline graphicmM to Inline graphicmM. In the model respiratory acidosis was caused by an increase in Inline graphic flux that cannot be vented from the extracellular space. Although it is recognised that the decrease in Inline graphic due to metabolic acidosis will also contribute to elevated Inline graphic, initially the Inline graphic flux was set at a Inline graphic level to achieve an increase in Inline graphic concentration to Inline graphicmM, which resulted in a decrease of Inline graphic to Inline graphic. During ischemia Inline graphic decreases rapidly before plateauing after approximately Inline graphic minutes. The decrease in the initial Inline graphic minutes has been reported to fall between Inline graphic to Inline graphic pH units in rat [108], ferret, [109] and guinea pig [110] preparations. Respiratory and metabolic acidosis will both contribute to this drop in Inline graphic. In the model setting the level of intracellular Inline graphic flux to Inline graphic caused a Inline graphic pH unit drop and an increase in Inline graphic to Inline graphicmM. Combined metabolic and respiratory acidosis caused a Inline graphic unit drop in pH and an increase in Inline graphic to Inline graphicmM.

In studies of ischemia Inline graphic tends to increase linearly with time. In the rat heart ischemia caused an increase in Inline graphic by Inline graphic over Inline graphic minutes [111], by Inline graphic after Inline graphic minutes [112], from Inline graphicmM to Inline graphicmM during Inline graphic minutes ischemia [113], by Inline graphic, Inline graphic, Inline graphic and Inline graphic after 9, Inline graphic, Inline graphic and Inline graphic minutes, respectively [114], by two fold over Inline graphic minutes [115], from Inline graphicmM to Inline graphicmM over Inline graphic minutes [116], by Inline graphic over Inline graphic minutes [117], by Inline graphic over Inline graphic minutes [118], no change over Inline graphic minutes [108] and by Inline graphic over Inline graphic minutes [119]. Changes in Inline graphic in guinea pig hearts during ischemia is controversial with reports of a decrease from Inline graphicmM to Inline graphicmM over Inline graphicminutes [120] and of an increase of Inline graphic over Inline graphic minutes [121]. Only considering the cases where Inline graphic increases, as these represent a repeatable consensus result, the range of expected increases in Inline graphic over a Inline graphic minute period is Inline graphic, assuming a linear increase with time. These cluster into two groups with ranges Inline graphic and Inline graphic. In the model reducing the maximum flux of Inline graphic by Inline graphic caused an increase in Inline graphic to Inline graphicmM (Inline graphic) in the absence of acidosis or Inline graphicmM (Inline graphic) in the presence of the acidotic components of ischemia, described above.

Given the inhibition of Inline graphic and levels of respiratory and metabolic acidosis a sweep of Inline graphic activation values was performed in the presence of these changes to achieve the desired level of extracellular Inline graphic accumulation. During ischemia, Inline graphic increases over three characteristic phases, with the first phase occurring during the initial Inline graphic minutes of ischemia prior to reaching a plateau from minutes Inline graphic to Inline graphic, before increasing again. In guinea pigs, ischemia caused Inline graphic to increase from Inline graphic to Inline graphicmM over Inline graphic minutes [122] or Inline graphic to Inline graphicmM over Inline graphic minutes [123]. In swine, ischemia caused Inline graphic to increase from Inline graphicmM to Inline graphicmM after Inline graphic minutes [124], [125]. In rat, ischemia caused an increase from Inline graphicmM to 8mM over Inline graphic minutes [110], although other groups have seen a biphasic change in Inline graphic in rat with an increase from Inline graphicmM to Inline graphicmM before falling back to Inline graphicmM then continuing to rise again, observed over the first Inline graphic minutes [126]. In rabbit, ischemia caused Inline graphic to increase from Inline graphicmM to Inline graphicmM over Inline graphic minutes [10], [110], [125]. In canine, Inline graphic increases from Inline graphicmM to Inline graphicmM after Inline graphic minutes [127], [128]. These results indicate an increase in Inline graphic from Inline graphicmM to Inline graphicmM after Inline graphic minutes and assuming Inline graphic of this rise occurs in the first Inline graphic minutes [127], and combined with the 5 minute data gives an estimate of Inline graphic after 5 minutes of ischemia as Inline graphicmM. In the model an increase in Inline graphic of Inline graphic was used to achieve an elevation of Inline graphic to Inline graphicmM, resulting in an increase in the membrane potential of Inline graphicmV to Inline graphicmV. This is consistent with measurements in cat (Inline graphicmV over Inline graphic minutes) [129], [130], sheep (Inline graphicmV over Inline graphic minutes) [131], guinea pigs (Inline graphicmV over Inline graphic minutes) [132] and mice (Inline graphicmV over Inline graphic minutes) [133], but less than measurements in rabbit (Inline graphicmV over Inline graphic minutes [10]) and guinea pig (Inline graphicmV over Inline graphic minutes [120] or Inline graphicmV over Inline graphic minutes [134][136]). The broad variation in membrane potential changes was not unexpected given the range of changes in Inline graphic and Inline graphic reported and in simulations an intermediate value has been achieved.

The individual and combined effects of the four components of ischemia on the temporal evolution of the maximum change in ionic concentrations and the spatial concentration and potential distribution profile after Inline graphic minutes of simulated ischemia are shown in Fig. 3. As ischemia progressed, Inline graphic continued to increase. The early rise was attributed to Inline graphic inhibition; the later increases were due to metabolic and respiratory acidosis. The early rise in Inline graphic was significantly affected by Inline graphic inhibition with Inline graphic activation playing a greater role as ischemia progresses and the membrane potential diverges from the Inline graphic reversal potential. The decrease in Inline graphic was solely due to acidosis with no impact from Inline graphic activation or Inline graphic inhibition. Inline graphic elevation was contributed to principally by Inline graphic inhibition, while increased Inline graphic decreased Inline graphic. Ischemia caused an initial drop followed by a sustained rise in Inline graphic. Inline graphic activation caused an early hyperpolarisation of the membrane potential, which was subsequently countered by the depolarising effects of Inline graphic inhibition, with acidosis having a limited effect.

Figure 3. Evolution and profile of A-B) Inline graphic, C-D) Inline graphic, E-F) Inline graphic, G-H) Inline graphic and I-J) membrane potential across the BZ due to each component of ischemia.

Figure 3

Complete ischemia, Inline graphic inhibition, Inline graphic inhibition, respiratory acidosis and metabolic acidosis are represented by black, yellow, red, blue and purple lines, respectively. Column 1 shows the profile of ionic concentrations and membrane potential after 5 minutes and column 2 shows the evolution of the change in magnitude in ionic concentrations and potential across the border zone with time.

Figure 4 plots the width of the BZ for each ion with striped bars corresponding to extracellular space and the darker the bar the more significant the concentration gradient relative to the initial concentration. This plot shows that Inline graphic had a significantly wider BZ with greater magnitude than Inline graphic. For pH regulation, Inline graphic had a narrower BZ compared to the significantly wider Inline graphic BZ, which may indicate the facilitation of proton transport via Inline graphic diffusion.

Figure 4. Width of ionic BZ. Gray scale represents magnitude of gradient with white indicating no gradient.

Figure 4

Solid bars indicate intracellular gradient and striped bars indicate extracellular gradient.

Extracellular Potassium Gradients

The Inline graphic BZ width was significantly wider than other BZ ion widths and notably larger than the Inline graphic BZ width. To determine the cause of this extended Inline graphic BZ the source of the cumulative changes in Inline graphic due to transmembrane flux, drift or diffusion over the 5 minutes of ischemia were calculated and plotted in Fig. 5. This showed that only the transmembrane flux had a significant gradient across the ischemic region. Separating the transmembrane flux into the Inline graphic and Inline graphic components then identified the Inline graphic channel, which includes the ATP-inactivated Inline graphic current, as the cause of this gradient. The gradient of IKb was due to the extensive membrane potential gradient into the ischemic region (see Fig. 3). To confirm that this was the cause of the Inline graphic BZ width, the membrane potential was calculated as normal, but an additional clamped membrane potential was calculated at each time step. The clamped membrane potential had the same maximum and minimum values as the correct membrane potential but instead of a smooth gradient across the BZ it had a sharp transition over the BZ. A comparison of this clamped and the control membrane potentials is shown in the Fig. 5C. The effects of using a clamped membrane potential to calculate the Inline graphic flux or the Inline graphic current on the cumulative changes in Inline graphic after 5 minutes of simulated ischemia are plotted in Fig. 5B, demonstrating that by removing the gradient in the membrane potential experienced by Inline graphic there is a significant narrowing in the Inline graphic BZ.

Figure 5. Mechanisms underpinning .

Figure 5

Inline graphic BZ width. A) Reference change in Inline graphic after 5 minutes of simulated ischemia caused by transmembrane flux (blue dashed line), diffusion (purple line), drift (red line) and the total change due to all causes (yellow line). The BZ is indicated by the yellow shaded region. B) Change in Inline graphic in the reference model (yellow line) compared with change in Inline graphic when Inline graphic (purple dashed line) or Inline graphic (blue dashed line) are exposed to a clamped membrane potential. The reference BZ and BZ when Inline graphic is exposed to a clamped membrane potential are indicated by the yellow and purple shaded regions, respectively.

Drift and Diffusion

To investigate the relative contribution of drift and diffusion to intra region fluxes (inter or extracellular) the drift and diffusion fluxes in each region were plotted, alongside the transmembrane flux, over the length of the strand after Inline graphic minutes of ischemia. Fig. 6 shows the differences in drift and diffusion between Inline graphic, Inline graphic and Inline graphic. As expected from the intra and extracellular gradients, intracellular Inline graphic and Inline graphic diffused into and out of the ischemic region in the intra and extracellular space, respectively. The converse was the case for Inline graphic. Due to the decrease in transmembrane potential, characteristic of ischemic regions, there was a convergence of intra and extracellular potentials with the extracellular potential decreasing in the ischemic region and the intracellular potential increasing. This gradient caused positive ions to drift into the ischemic region in the extracellular space and drift out of the ischemic region in the intracellular space. The converse was true for negatively charged ions. As drift is proportional to the ionic concentration, Inline graphic drift was significant in the extracellular space and Inline graphic drift was significant in the intracellular space. For Inline graphic, drift and diffusion operated in the same direction in the intra and extracellular space. The result was a cyclical movement of Inline graphic moving into the ischemic region in the extracellular space, while moving out of the ischemic region in the intracellular space. The Inline graphic movement was also circular but in the opposite direction (Fig. 6P). However, for Inline graphic drift and diffusion were in opposite directions. In the extracellular space where the Inline graphic concentration was low, diffusion dominated and Inline graphic moved out of the ischemic region. In the intracellular space, where there was a higher concentration of Inline graphic, drift dominated, also causing Inline graphic ions to move out of the ischemic region. Thus ischemia caused a depletion of Inline graphic in the ischemic region through both the intra and extracellular space and, contrary to previous hypothesis [13], the model suggests that intracellular Inline graphic movement is the dominant path for Inline graphic to leave the ischemic region.

Figure 6. Regulation of Inline graphic, Inline graphic and Inline graphic ionic concentrations across the border zone after 5 minutes of ischemia.

Figure 6

Intra and extracellular A) Inline graphic, B) Inline graphic and C) Inline graphic ionic concentrations. D) Intra and extracellular potential. Intracellular drift and diffusion flux of E) Inline graphic, F) Inline graphic and G) Inline graphic. Extracellular drift and diffusion of H) Inline graphic, I) Inline graphic and J) Inline graphic. Transmembrane flux of K) Inline graphic, L) Inline graphic and M) Inline graphic. Schematics showing the general direction of ion movement within and between the intra and extracellular space for N) Inline graphic, O) Inline graphic and P) Inline graphic.

Discussion

In this study a new model of cardiac tissue electrophysiology was developed. The model predicted that the width of the Inline graphic gradient across an ischemic BZ would be significantly wider the Inline graphic BZ. The cause of this difference was attributed to the voltage dependence of the Inline graphic channel. The model also demonstrated that, due to electrogenic drift, Inline graphic moved out of the ischemic region in both the intra and extracellular space which will lead to Inline graphic depletion.

The model of ionic movement and tissue electrophysiology was developed by combining the Nernst-Plank equations with the bidomain framework. No attempt was made to explicitly validate the proposed tissue model equations due to the paucity of experimental data. However, applying simplifying assumptions with regards to ionic or voltage gradients reduces the proposed equations to the well validated bidomain equations [137], [138] or coupled reaction-diffusion equations [28], [86], respectively, providing support for the validity of the proposed modelling framework. The limited attempts at simulating the spatial temporal evolution of ionic gradients across the ischemic BZ have largely uncoupled the movement of ions and the electric field. Potse et al., [18] demonstrated that measured Inline graphic gradients across an ischemic BZ could be simulated using a model of Inline graphic diffusion coupled to a source term. The spatially varying Inline graphic gradient could then be included as a boundary condition to models of transmembrane current in the bidomain equations. This model did not include any effect of electric gradients on Inline graphic movement, Inline graphic movement, inter ionic species interactions or the effect of the ischemic region on any other ion gradient. Similar sets of equations to those proposed here have been used for simulating the potential gradient surrounding cells, including the Debye layer [139], electrical propagation along strands of cardiac cells [140], [141] and for modelling ion diffusion in the cable equation [142]. These previous models have either explicitly represented the intra and extracellular domains or only considered the intracellular domain but have not modelled the tissue within the bidomain framework, as derived and implemented here.

The proposed equations can be applied generally in three dimensions as opposed to the one dimensional simulations presented here. The current study does not consider the effects of anisotropy on ion or membrane potential gradients across the BZ, however, if implemented in two or three dimensions the model is capable of representing tissue anisotropy and any effects this may have on BZ gradients. By conserving ionic species the proposed equations provide a more biophysical representation of cardiac electrophysiology than the bidomain equations and can appropriately be applied to simulate a broader range of conditions. However, these benefits come at a cost. Unlike the bidomain equations, with two partial differential equations that can readily be uncoupled and solved as two sets of linear equations [143], the proposed model is nonlinear and contains two parabolic partial differential equations for each ionic species and one elliptic partial differential equation to model the electric potential. This results in a significant increase both in the complexity and number of equations that must be solved and hence comes at a significant increase in computational cost. The proposed framework can be used to simulate electrically active tissue by introducing a full action potential cell model [31]. This would require identifying and separating out the transmembrane pathways for each ionic species present in the cell model to calculate the net transmebrane flux for each ion or Inline graphic in Eqn 18. The internal cell model state variables, including gating variables, Markov states and intracellular Inline graphic dynamics would also need to be solved, introducing a system of nonlinear ordinary differential equations at each grid point. The small time steps and increased number of degrees of freedom required to simulate electrically active tissue would further increase the computational cost of the proposed model.

A new model of ion movement across the un-stimulated cell membrane was developed. This model has been published online and is available at cellml.org. The model creation approach demonstrated two novel methods. Firstly, the ion transporter densities were uniquely constrained by a small number of experimental data sets and a zero net flux constraint. This provided a repeatable and unique method for determining model parameters under quiescent conditions and could readily be applied to any cardiac electrophysiology model to constrain model parameters. Secondly, in the development of this model a grid of Inline graphic experimental observations were created to provide a comprehensive validation of the model response to changes in ion concentrations and in the presence of blockers of major transporters. Although these observations only provide a qualitative comparison, they do provide a benchmark for quantifying the generality of models. Furthermore, Table 4 identified Inline graphic experiments that do not appear to have been performed or remain controversial, highlighting the potential for additional experiments.

In this study the process of validating the model against Inline graphic experimental observations demonstrated the general capacity of the model to replicate the majority of experimental results. However, the model was unable to match ten of the Inline graphic observations. Five of the experimental observations found no change in a measurement. Due to the numerical nature of the model, even very small changes in a concentration can be observed and without adding in a semi arbitrary threshold that should reflect the variability and confidence of each experimental measurement it was not possible for the model to return no change in a value. In general the five remaining failed observations can be attributed to absent mechanisms in the model or inconsistencies with the model and specific experimental setups. The model did not predict the Inline graphic response to depressed Inline graphic and the Inline graphic response to depressed Inline graphic. This is potentially due to the absence of NaK2Cl from the model, largely due to the lack of data characterising the transporter. The model was unable to replicate the decrease in Inline graphic due to elevated Inline graphic. This may be due to the specifics of the experimental protocol. Decreasing Inline graphic, which should have the opposite effect on Inline graphic, caused both increases and decreases in Inline graphic in different studies indicating that the Inline graphic dependence on Inline graphic is sensitive to the specifics of the experimental setup. The model did not predict the change in Inline graphic or Inline graphic with a decrease in Inline graphic. Experimental observations report an increase in Inline graphic with both a decrease or an increase in Inline graphic. This may indicate that the Inline graphic is at some minima with respect to Inline graphic, although this seems unlikely. It more likely reflects inconsistent data due to differences in experimental setups that were impossible for the model to replicate. In cardiac myocytes it is postulated that Inline graphic and Inline graphic compete for common buffering sites [144], [145]. The decrease in Inline graphic may drain the cell of Inline graphic, reducing the Inline graphic bound to buffers that could then be occupied by Inline graphic, resulting in a decrease in Inline graphic. A common pool of Inline graphic and Inline graphic buffering was not present in the model and this may explain the disparity between model predictions and experimental results. Despite the inability of the model to replicate Inline graphic of the Inline graphic observations, the majority of the absent mechanisms are expected to play a secondary role during ischemia. NaK2Cl is a potential contributor to the elevation in Inline graphic [146], although this has been questioned [131], and its absence is unlikely to affect the general model conclusions.

Models simulations found a BZ width for the different ions of between Inline graphicmm (Inline graphic) and Inline graphicmm (Inline graphic). The variation in the width of each ion across the BZ means that there was no single BZ width predicted by the model. The majority of ions transitioned from viable to ischemic concentrations over Inline graphicmm, with Inline graphic and Inline graphic notable outliers. The model predictions are consistent with previous measurements of the BZ between Inline graphic and Inline graphicmm [10], [11], [147][149]. However, the model did not include the sharp gradients in metabolites over Inline graphicmm [11] or the effects of regions of reduced perfusion affecting mechanics over 20 mm distance from the ischemic region [150].

The model predicted that Inline graphic and Inline graphic have significantly different BZ widths (Fig. 4). It was expected that the coupling of Inline graphic and Inline graphic via Inline graphic would result in concordant gradients in these two ions across the BZ. The extended Inline graphic gradient was attributed to the voltage dependence of the Inline graphic (Fig 5). Unlike Inline graphic, Inline graphic is regulated by two electrogenic transmembrane pathways Inline graphic and Inline graphic, whereas Inline graphic is regulated by multiple exchangers that are electro neutral, or have attenuated voltage dependence, compared to ion channels. This discordance in Inline graphic and Inline graphic gradients will have a significant effect on the gradient of action potential morphology across the BZ. Elevation of Inline graphic depolarises the cell and shortens the action potential duration, whereas elevation of Inline graphic causes a reduction in action potential duration [151], [152]. The combined effect of the two gradients will be a slower change in resting membrane potential on the length scale of the Inline graphic gradient and a much more rapid transition in action potential duration due to the faster change in Inline graphic in conjunction with the change in Inline graphic over the BZ. The temporal evolution of the two gradients are also distinct, with a sustained constant increase in Inline graphic over the first Inline graphic minutes, while Inline graphic increases rapidly for Inline graphic minutes before plateauing (Fig. 3). These simulation results are consistent with experimental measurements of ionic concentrations [111], [152] and changes in ECG morphology, which report an early elevation of resting membrane potential, followed by a decrease in action potential duration [153] during early ischemia. These spatial-temporal increases in ion concentrations and secondary effects on electrophysiology will increase tissue heterogeneity and have the capacity to play an important role in the BZ arrhythmogenic substrate.

Experimental measurements have reported a decrease in Inline graphic during ischemia [152]. Loss of Inline graphic has been attributed to a combination of increased Inline graphic conduction and Inline graphic inhibition [152], [154]. However, depletion of Inline graphic due to transmembrane Inline graphic movement, in the absence of a potential gradient, would cause an intracellular flux of Inline graphic into the ischemic zone, due to diffusion, to replenish Inline graphic, mitigating the effects of changes of Inline graphic transmembrane flux on Inline graphic. The model proposed here demonstrated that this is not necessarily the case. The model predicted a significant electrogenic Inline graphic flux in the intracellular space out of the ischemic zone (Fig. 6), resulting in a net efflux of Inline graphic out of the ischemic zone both in the intra and extracellular space. The movement of Inline graphic in the intracellular space out of the ischemic zone would further deplete Inline graphic in the ischemic region and exacerbate Inline graphic loss, but may limit Inline graphic accumulation in the extracellular space.

Limitations

The model is inherently an approximation and hence represents a finite set of known cellular properties and changes that occur during ischemia. In particular, the model treated all buffers as static and rapid, the effects of protons on channel, exchanger and co-transports were not considered, ischemic changes were instantaneous and the model did not include all possible changes or pathways that may affect ionic homeostasis during ischemia.

The tissue model assumed that Inline graphic and Inline graphic buffers are static, rapid and made up of a single population of binding sites. It is known that some of the buffers for both Inline graphic and Inline graphic are mobile, these could be introduced into the model framework as an additional concentration but these effects were approximated, without the additional computational cost of adding a additional ionic concentration, by using effective diffusion constants. The equilibrium assumption was likely to be valid in the current model due to the long time scales of interest; however, simulation of cardiac action potentials would require the re-evaluation of this assumption. It is also known that Inline graphic [28] and Inline graphic [155] are buffered by multiple proteins with distinct binding kinetics but over the range of concentrations simulated these multiple buffer species were unlikely to have significant effects.

The model of intracellular Inline graphic dynamics assumes that the SR and cytosolic Inline graphic concentrations remain in equilibrium, which is clearly not the case during an action potential. The model of intracellular Inline graphic dynamics provided a numerically efficient representation of intracellular Inline graphic buffering and SR Inline graphic uptake. The use of a Hill coefficient of one for SERCA as opposed to the more common and biophysical value of two removed the need for an additional differential equation to model Inline graphic or the solution of a set of nonlinear equations to model Inline graphic regulation. Furthermore, over the range of Inline graphic values studied it was possible to adjust the maximum SERCA flux to minimize discrepancies between a model with a Hill coefficient of one or two.

It is well recognised that Inline graphic play an important role in regulating cellular electrophysiology [156]. Experimental and modelling studies have demonstrated the effects of Inline graphic on ryanodine receptor opening probabilities, Inline graphic, Inline graphic channels, Inline graphic buffering and SERCA [22], [157]. The majority of the effects of Inline graphic on Inline graphic regulation are unlikely to play a significant role in determining the spatial and temporal Inline graphic and Inline graphic. However, inhibition of Inline graphic potentially contributes to Inline graphic gradients but this will be secondary to the effects of Inline graphic inhibition.

The model treats all changes for ischemia as instantaneous. The time dependence of inhibition of Inline graphic, activation of Inline graphic, increased Inline graphic flux or increased Inline graphic flux are not known and can only be approximated. A linear ramp in ischemic changes has been used previously, but this fails to consider the possibility that changes occur over different time scales. In order to minimise ambiguity in model simulations an instantaneous change in Inline graphic, Inline graphic, Inline graphic flux and Inline graphic flux was chosen.

To limit the scope of this study, the effects of cell swelling and the effects of this on changes in ion concentrations [122] were not included in the model. However, previous, modelling studies have found these effects to not significantly alter Inline graphic accumulation and may not fundamentally alter the study conclusions [20]. Mitani and Shattock [146] identified the Na dependent potassium channel as a potential contributor to the elevation of Inline graphic. However, the effects of an increase in any Inline graphic current are captured by the elevation in Inline graphic that does not necessarily need to be the sole result of an increase in conduction in the ATP inactivated Inline graphic channel. Furthermore, the model does not include an increased Inline graphic channel conductance during ischemia. This has been reported during ischemia in the form of increased permeability of Inline graphic through the persistent Inline graphic current [116] and the ATP activated Inline graphic channel [158]. However, other groups have found a limited impact of the persistent Inline graphic channel in Inline graphic accumulation during ischemia [159] and previous modelling studies have shown that its inclusion is not required to capture the salient features of ischemia in the single cell [20]. For these reasons a potential increase in the Inline graphic channel conductance was not included in this study.

Summary

A new mathematical framework was derived for simulating cardiac tissue electrophysiology with ion species conservation. The model was used to simulate the movement of ions due to transmembrane channels, pumps and transporters, diffusion and drift in the intra and extracellular space. The model predicted that 1) the sodium BZ is approximately a quarter of the length of the potassium BZ and this is due to the effects of the membrane potential gradient on Inline graphic and 2) that during ischemia there is a gross movement of potassium ions out of the ischemic region in both the intra and extracellular space due to the effects of drift, which will lead to a depletion of Inline graphic from the ischemic region.

Funding Statement

The work was supported by UK Engineering and Physical Sciences Research Council EP/F043929/1, British Heart Foundation (PG/11/101/29212) and Boston Scientific. The funders had no role in study design, data collection and analysis, decision to publish, or preparation of the manuscript.

References

  • 1. Kaplinsky E, Ogawa S, Balke CW, Dreifus LS (1979) Two periods of early ventricular arrhythmia in the canine acute myocardial infarction model. Circulation 60: 397–403. [DOI] [PubMed] [Google Scholar]
  • 2. Janse MJ, Cinca J, Morena H, Fiolet JW, Kleber AG, et al. (1979) The “border zone” in myocardial ischemia. an electrophysiological, metabolic, and histochemical correlation in the pig heart. Circulation Research 44: 576–588. [DOI] [PubMed] [Google Scholar]
  • 3. Cascio WE (2001) Myocardial ischemia: What factors determine arrhythmogenesis? Journal of Cardiovascular Electrophysiology 12: 726–729. [DOI] [PubMed] [Google Scholar]
  • 4. Zaitsev AV, Guha PK, Sarmast F, Kolli A, Berenfeld O, et al. (2003) Wavebreak formation during ventricular fibrillation in the isolated, regionally ischemic pig heart. Circulation Research 92: 546–553. [DOI] [PubMed] [Google Scholar]
  • 5. Carmeliet E (1999) Cardiac ionic currents and acute ischemia: From channels to arrhythmias. physiological reviews 73: 917–1017. [DOI] [PubMed] [Google Scholar]
  • 6. Sidorov VY, Uzelac I, Wikswo JP (2011) Regional increase of extracellular potassium leads to electrical instability and reentry occurrence through the spatial heterogeneity of apd restitution. American Journal of Physiology - Heart and Circulatory Physiology 301: H209–H220. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7. Weiss JN, Chen PS, Qu Z, Karagueuzian HS, Garfinkel A (2000) Ventricular fibrillation. Circulation Research 87: 1103–1107. [DOI] [PubMed] [Google Scholar]
  • 8. Horacek T, Neumann M, von Mutius S, Budden M, Meesmann W (1984) Nonhomogeneous electrophysiological changes and the bimodal distribution of early ventricular arrhythmias during acute coronary artery occlusion. Basic Research in Cardiology 79: 649–667. [DOI] [PubMed] [Google Scholar]
  • 9. Klber AG (1984) Extracellular potassium accumulation in acute myocardial ischemia. Journal of Molecular and Cellular Cardiology 16: 389–394. [DOI] [PubMed] [Google Scholar]
  • 10. Wilensky RL, Tranum-Jensen J, Coronel R, Wilde AA, Fiolet JW, et al. (1986) The subendocardial border zone during acute ischemia of the rabbit heart: an electrophysiologic, metabolic, and morphologic correlative study. Circulation 74: 1137–1146. [DOI] [PubMed] [Google Scholar]
  • 11. Harken AH, Barlow CH, Harden Iii WR, Chance B (1978) Two and three dimensional display of myocardial ischemic border zone in dogs. The American Journal of Cardiology 42: 954–959. [DOI] [PubMed] [Google Scholar]
  • 12. Yellon DM, Hearse DJ, Crome R, Grannell J, Wyse RKH (1981) Characterization of the lateral interface between normal and ischemic tissue in the canine heart during evolving myocardial infarction. The American Journal of Cardiology 47: 1233–1239. [DOI] [PubMed] [Google Scholar]
  • 13. Coronel R, Fiolet JW, Wilms-Schopman FJ, Schaapherder AF, Johnson TA, et al. (1988) Distribution of extracellular potassium and its relation to electrophysiologic changes during acute myocardial ischemia in the isolated perfused porcine heart. Circulation 77: 1125–1138. [DOI] [PubMed] [Google Scholar]
  • 14. Coronel R, Wilms-Schopman FJG, Fiolet JWT, Opthof T, Janse MJ (1995) The relation between extracellular potassium concentration and ph in the border zone during regional ischemia in isolated porcine hearts. Journal of Molecular and Cellular Cardiology 27: 2069–2073. [DOI] [PubMed] [Google Scholar]
  • 15. Tice BM, Rodrguez B, Eason J, Trayanova N (2007) Mechanistic investigation into the arrhythmogenic role of transmural heterogeneities in regional ischaemia phase 1a. Europace 9: vi46–vi58. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16. Trnor B, Ferrero JM, Rodrguez B, Montilla F (2005) Effects of pinacidil on reentrant arrhythmias generated during acute regional ischemia: A simulation study. Annals of Biomedical Engineering 33: 897–906. [DOI] [PubMed] [Google Scholar]
  • 17. Potse M, Coronel R, Falcao S, LeBlanc AR, Vinet A (2007) The effect of lesion size and tissue remodeling on ST deviation in partial-thickness ischemia. Heart Rhythm 4: 200–206. [DOI] [PubMed] [Google Scholar]
  • 18. Potse M, Coronel R, LeBlanc AR, Vinet A (2007) The role of extracellular potassium transport in computer models of the ischemic zone. Medical and Biological Engineering and Computing 45: 1187–1199. [DOI] [PubMed] [Google Scholar]
  • 19. Smith NP, Crampin EJ (2004) Development of models of active ion transport for whole-cell modelling: cardiac sodium-potassium pump as a case study. Progress in Biophysics & Molecular Biology 85: 387–405. [DOI] [PubMed] [Google Scholar]
  • 20. Terkildsen JR, Crampin EJ, Smith NP (2007) The balance between inactivation and activation of the Na+-K+ pump underlies the triphasic accumulation of extracellular K+ during myocardial ischemia. Am J Physiol Heart Circ Physiol 293: H3036–3045. [DOI] [PubMed] [Google Scholar]
  • 21. Weber CR, Piacentino I Valentino, Ginsburg KS, Houser SR, Bers DM (2002) Na+-Ca2+ exchange current and submembrane [Ca2+]. during the cardiac action potential. Circulation Research 90: 182–189. [DOI] [PubMed] [Google Scholar]
  • 22. Crampin EJ, Smith NP (2006) A dynamic model of excitation-contraction coupling during acidosis in cardiac ventricular myocytes. Biophysical Journal 90: 3074–3090. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23. Niederer SA, Smith NP (2007) A mathematical model of the slow force response to stretch in rat ventricular myocytes. Biophysical Journal 92: 4030–44. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24. Wu ML, Vaughan-Jones RD (1997) Interaction between Na+ and H+ ions on Na-H exchange in sheep cardiac purkinje fibers. Journal of Molecular and Cellular Cardiology 29: 1131–1140. [DOI] [PubMed] [Google Scholar]
  • 25. Zaniboni M, Swietach P, Rossini A, Yamamoto T, Spitzer KW, et al. (2003) Intracellular proton mobility and buffering power in cardiac ventricular myocytes from rat, rabbit, and guinea pig. American Journal of Physiology 285: H1236–H1246. [DOI] [PubMed] [Google Scholar]
  • 26. Despa S, Islam MA, Pogwizd SM, Bers DM (2002) Intracellular Na+. and Na+ pump rate in rat and rabbit ventricular myocytes. J Physiol (Lond) 539: 133–143. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27. Niederer SA, Swietach P, Wilson DA, Smith NP, Vaughan-Jones RD (2008) Measuring and modeling chloride-hydroxyl exchange in the guinea-pig ventricular myocyte. Biophysical Journal 94: 2385–403. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28. Zaniboni M, Rossini A, Swietach P, Banger N, Spitzer KW, et al. (2003) Proton permeation through the myocardial gap junction. Circulation Research 93: 726–735. [DOI] [PubMed] [Google Scholar]
  • 29. Faber GM, Rudy Y (2000) Action potential and contractility changes in Na+ (i) overloaded cardiac myocytes: A simulation study. Biophysical Journal 78: 2392–2404. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30. Komukai K, Brette F, Pascarel C, Orchard CH (2002) Electrophysiological response of rat ventricular myocytes to acidosis. American Journal of Physiology - Heart and Circulatory Physiology 283: H412–H422. [DOI] [PubMed] [Google Scholar]
  • 31. Fink M, Niederer SA, Cherry EM, Fenton FH, Koivumki JT, et al. (2011) Cardiac cell modelling: Observations from the heart of the cardiac physiome project. Progress in Biophysics and Molecular Biology 104: 2–21. [DOI] [PubMed] [Google Scholar]
  • 32. Monnereau-Soustre H, Revial MF, Leoty C, Pager J (1969) evaluation of extracellular space of the rat heart: effect of adrenalectomy; electrophysiological implications. C R Seances Soc Biol Fil 163: 1419–22. [PubMed] [Google Scholar]
  • 33. Polimeni P (1974) Extracellular space and ionic distribution in rat ventricle. American Journal of Physiology – Legacy Content 227: 676–683. [DOI] [PubMed] [Google Scholar]
  • 34. Poole-Wilson P, Cameron I (1975) ECS, intracellular pH, and electrolytes of cardiac and skeletal muscle. American Journal of Physiology – Legacy Content 229: 1299–1304. [DOI] [PubMed] [Google Scholar]
  • 35. Caille JP, Ruiz-Ceretti E, Schanne OF (1981) Intracellular chloride activity in rabbit papillary muscle: effect of ouabain. American Journal of Physiology - Cell Physiology 240: C183–C188. [DOI] [PubMed] [Google Scholar]
  • 36. Polimeni PI (1974) Cardiac electrolytes and water in thyroparathyroidectomized rat. Journal of Molecular and Cellular Cardiology 6: 531–541. [DOI] [PubMed] [Google Scholar]
  • 37. Page E (1978) Quantitative ultrastructural analysis in cardiac membrane physiology. American Journal of Physiology - Cell Physiology 235: C147–C158. [DOI] [PubMed] [Google Scholar]
  • 38. Page E, Surdyk-Droske M (1979) Distribution, surface density, and membrane area of diadic junctional contacts between plasma membrane and terminal cisterns in mammalian ventricle. Circulation Research 45: 260–267. [DOI] [PubMed] [Google Scholar]
  • 39. Chopra N, Kannankeril PJ, Yang T, Hlaing T, Holinstat I, et al. (2007) Modest reductions of cardiac calsequestrin increase sarcoplasmic reticulum Ca2+ leak independent of luminal Ca2+ and trigger ventricular arrhythmias in mice. Circulation Research 101: 617–626. [DOI] [PubMed] [Google Scholar]
  • 40. Singh S, White F, Bloor C (1981) Myocardial morphometric characteristics in swine. Circulation Research 49: 434–441. [DOI] [PubMed] [Google Scholar]
  • 41. Abete P, Vassalle M (1988) Relation between Na+-K+ pump, Na+ activity and force in strophanthidin inotropy in sheep cardiac purkinje fibres. The Journal of Physiology 404: 275–299. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42. Sossalla S, Wagner S, Rasenack ECL, Ruff H, Weber SL, et al. (2008) Ranolazine improves diastolic dysfunction in isolated myocardium from failing human hearts – role of late sodium current and intracellular ion accumulation. Journal of Molecular and Cellular Cardiology 45: 32–43. [DOI] [PubMed] [Google Scholar]
  • 43. Kimura S, Bassett AL, Gaide MS, Kozlovskis PL, Myerburg RJ (1986) Regional changes in intracellular potassium and sodium activity after healing of experimental myocardial infarction in cats. Circulation Research 58: 202–208. [DOI] [PubMed] [Google Scholar]
  • 44. Baartscheer A, Hardziyenka M, Schumacher CA, Belterman CNW, Borren MMGJv, et al. (2008) Chronic inhibition of the Na-H exchanger causes regression of hypertrophy, heart failure, and ionic and electrophysiological remodelling. British Journal of Pharmacology 154: 1266–1275. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45. Lee CO, Fozzard HA (1975) Activities of potassium and sodium ions in rabbit heart muscle. The Journal of General Physiology 65: 695–708. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46. Prasad K, Khatter JC, Bharadwaj B (1979) Intra- and extracellular electrolytes and sarcolemmal ATPase in the failing heart due to pressure overload in dogs. Cardiovascular Research 13: 95–104. [DOI] [PubMed] [Google Scholar]
  • 47. Cohen C, Fozzard H, Sheu S (1982) Increase in intracellular sodium ion activity during stimulation in mammalian cardiac muscle. Circulation Research 50: 651–662. [DOI] [PubMed] [Google Scholar]
  • 48. Baumgarten CM, Fozzard HA (1981) Intracellular chloride activity in mammalian ventricular muscle. American Journal of Physiology - Cell Physiology 241: C121–C129. [DOI] [PubMed] [Google Scholar]
  • 49. Powell T, Terrar DA, Twist VW (1980) Electrical properties of individual cells isolated from adult rat ventricular myocardium. The Journal of Physiology 302: 131–153. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50. Vaughan-Jones RD (1979) Non-passive chloride distribution in mammalian heart muscle: micro-electrode measurement of the intracellular chloride activity. The Journal of Physiology 295: 83–109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51. Lai ZF, Nishi K (1998) Intracellular chloride activity increases in guinea pig ventricular muscle during simulated ischemia. American Journal of Physiology - Heart and Circulatory Physiology 275: H1613–H1619. [DOI] [PubMed] [Google Scholar]
  • 52.Bers DM (2001) excitation-contraction coupling and cardiac contractile force. Dordrecht: Kluwer Academic Publishers, first edition.
  • 53. Baartscheer A, Schumacher CA, Belterman CN, Coronel R, Fiolet JW (2003) SR calcium handling and calcium after-transients in a rabbit model of heart failure. Cardiovascular Research 58: 99–108. [DOI] [PubMed] [Google Scholar]
  • 54. Bassani JW, Bassani RA, Bers DM (1995) Calibration of indo-1 and resting intracellular [Ca]i in intact rabbit cardiac myocytes. Biophysical Journal 68: 1453–1460. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55. Siri FM, Krueger J, Nordin C, Ming Z, Aronson RS (1991) Depressed intracellular calcium transients and contraction in myocytes from hypertrophied and failing guinea pig hearts. American Journal of Physiology - Heart and Circulatory Physiology 261: H514–H530. [DOI] [PubMed] [Google Scholar]
  • 56. Han S (1994) Schiefer A, Isenberg G (1994) Ca2+ load of guinea-pig ventricular myocytes determines efficacy of brief Ca2+ currents as trigger for Ca2+ release. The Journal of Physiology 480: 411–421. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57. Nakanishi T, Seguchi M, Tsuchiya T, Yasukouchi S, Takao A (1990) Effect of acidosis on intracellular ph and calcium concentration in the newborn and adult rabbit myocardium. Circulation Research 67: 111–123. [DOI] [PubMed] [Google Scholar]
  • 58. Terracciano CM, MacLeod KT (1997) Measurements of Ca2+ entry and sarcoplasmic reticulum Ca2+ content during the cardiac cycle in guinea pig and rat ventricular myocytes. Biophysical journal 72: 1319–1326. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59. Varro A, Negretti N, Hester SB, Eisner DA (1993) An estimate of the calcium content of the sarcoplasmic reticulum in rat ventricular myocytes. Pflügers Archiv European Journal of Physiology 423: 158–160. [DOI] [PubMed] [Google Scholar]
  • 60. Bassani RA, Bers DM (1995) Rate of diastolic Ca release from the sarcoplasmic reticulum of intact rabbit and rat ventricular myocytes. Biophysical Journal 68: 2015–2022. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61. Delbridge LM, Satoh H, Yuan W, Bassani JW, Qi M, et al. (1997) Cardiac myocyte volume, Ca2+ fluxes, and sarcoplasmic reticulum loading in pressure-overload hypertrophy. American Journal of Physiology - Heart and Circulatory Physiology 272: H2425–H2435. [DOI] [PubMed] [Google Scholar]
  • 62. Hobai IA, ORourke B (2001) Decreased sarcoplasmic reticulum calcium content is responsible for defective excitation-contraction coupling in canine heart failure. Circulation 103: 1577–1584. [DOI] [PubMed] [Google Scholar]
  • 63. Trafford AW, Daz ME, Eisner DA (1999) A novel, rapid and reversible method to measure ca buffering and time-course of total sarcoplasmic reticulum ca content in cardiac ventricular myocytes. Pflgers Archiv European Journal of Physiology 437: 501–503. [DOI] [PubMed] [Google Scholar]
  • 64. Bennett DL, ONeill SC, Eisner DA (1999) Strophanthidin-induced gain of Ca2+ occurs during diastole and not systole in guinea-pig ventricular myocytes. Pflugers Archiv European Journal of Physiology 437: 731–736. [DOI] [PubMed] [Google Scholar]
  • 65. Hove-Madsen L, Bers D (1993) Sarcoplasmic reticulum Ca2+ uptake and thapsigargin sensitivity in permeabilized rabbit and rat ventricular myocytes. Circulation Research 73: 820–828. [DOI] [PubMed] [Google Scholar]
  • 66. Berlin JR, Bassani JW, Bers DM (1994) Intrinsic cytosolic calcium buffering properties of single rat cardiac myocytes. Biophysical Journal 67: 1775–1787. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67. Walden AP, Dibb KM, Trafford AW (2009) Differences in intracellular calcium homeostasis between atrial and ventricular myocytes. Journal of Molecular and Cellular Cardiology 46: 463–473. [DOI] [PubMed] [Google Scholar]
  • 68. Díaz ME, Trafford AW, Eisner DA (2001) The effects of exogenous calcium buffers on the systolic calcium transient in rat ventricular myocytes. Biophysical Journal 80: 1915–1925. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69. Gottlieb RA, Gruol DL, Zhu JY, Engler RL (1996) Preconditioning rabbit cardiomyocytes: role of ph, vacuolar proton atpase, and apoptosis. J Clin Invest 97: 2391–8. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70. Leem CH, Lagadic-Gossmann D, Vaughan-Jones RD (1999) Characterization of intracellular pH regulation in the guinea-pig ventricular myocyte. Journal of Physiology 517: 159–180. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71. Lagadic-Gossmann D, Buckler KJ, Vaughan-Jones RD (1992) Role of bicarbonate in pH recovery from intracellular acidosis in the guinea-pig ventricular myocyte. Journal of Physiology 458: 361–384. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72. Cummings JR (1960) Electrolyte changes in heart tissue and coronary arterial and venous plasma following coronary occlusion. Circulation Research 8: 865–870. [DOI] [PubMed] [Google Scholar]
  • 73. Polimeni P, Al-Sadir J (1975) Expansion of extracellular space in the nonischemic zone of the infarcted heart and concomitant changes in tissue electrolyte contents in the rat. Circulation Research 37: 725–732. [DOI] [PubMed] [Google Scholar]
  • 74. Boyer PK, Poindexter CA (1940) The influence of digitalis on the electrolyte and water balance of heart muscle. American Heart Journal 20: 586–591. [Google Scholar]
  • 75. Erdmann E, Bolte HD, Lderitz B (1971) The (Na++K+)-atpase activity of guinea pig heart muscle in potassium deficiency. Archives of Biochemistry and Biophysics 145: 121–125. [DOI] [PubMed] [Google Scholar]
  • 76. Kunze D (1977) Rate-dependent changes in extracellular potassium in the rabbit atrium. Circulation Research 41: 122–127. [DOI] [PubMed] [Google Scholar]
  • 77. Moore EW (1970) Ionized calcium in normal serum, ultrafiltrates, and whole blood determined by ion-exchange electrodes. J Clin Invest 49: 318–34. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78. Post JA, Langer GA (1992) Sarcolemmal calcium binding sites in heart: I. molecular origin in “gas-dissected” sarcolemma. Journal of Membrane Biology 129: 49–57. [DOI] [PubMed] [Google Scholar]
  • 79. Bers DM, Philipson KD, Langer GA (1981) Cardiac contractility and sarcolemmal calcium binding in several cardiac muscle preparations. American Journal of Physiology - Heart and Circulatory Physiology 240: H576–H583. [DOI] [PubMed] [Google Scholar]
  • 80. Gayeski TE, Honig CR (1991) Intracellular PO2 in individual cardiac myocytes in dogs, cats, rabbits, ferrets, and rats. American Journal of Physiology - Heart and Circulatory Physiology 260: H522–H531. [DOI] [PubMed] [Google Scholar]
  • 81. Yan G, Kleber A (1992) Changes in extracellular and intracellular ph in ischemic rabbit papillary muscle. Circulation Research 71: 460–470. [DOI] [PubMed] [Google Scholar]
  • 82. Leem CH, Vaughan-Jones RD (1998) Sarcolemmal mechanisms for pH(i) recovery from alkalosis in the guinea-pig ventricular myocyte. Journal of Physiology 509: 487–496. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83. Austin WH, Lacombe E, Rand PW, Chatterjee M (1963) Solubility of carbon dioxide in serum from 15 to 38 c. Journal of Applied Physiology 18: 301–304. [DOI] [PubMed] [Google Scholar]
  • 84. Conway RS, Weiss HR (1985) Dependence of spatial heterogeneity of myocardial blood flow on mean blood flow rate in the rabbit heart. Cardiovascular Research 19: 160–168. [DOI] [PubMed] [Google Scholar]
  • 85. Despa S, Bers DM (2003) Na/K pump current and Na.i in rabbit ventricular myocytes: Local Na.i depletion and Na buffering. Biophysical Journal 84: 4157–4166. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86. Swietach P, Zaniboni M, Stewart AK, Rossini A, Spitzer KW, et al. (2003) Modelling intracellular H+ ion diffusion. Progress in Biophysics and Molecular Biology 83: 69–100. [DOI] [PubMed] [Google Scholar]
  • 87. Bassani JW, Bassani RA, Bers DM (1994) Relaxation in rabbit and rat cardiac cells: species-dependent differences in cellular mechanisms. The Journal of Physiology 476: 279–293. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88. Balke CW, Egan TM, Wier WG (1994) Processes that remove calcium from the cytoplasm during excitation-contraction coupling in intact rat heart cells. The Journal of Physiology 474: 447–462. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89. Shannon TR, Ginsburg KS, Bers DM (2002) Quantitative assessment of the SR Ca2+ leak-load relationship. Circulation Research 91: 594–600. [DOI] [PubMed] [Google Scholar]
  • 90. Zima AV, Bovo E, Bers DM, Blatter LA (2010) Ca2+ spark-dependent and -independent sarcoplasmic reticulum Ca2+ leak in normal and failing rabbit ventricular myocytes. The Journal of Physiology 588: 4743–4757. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91. Li L, Niederer SA, Idigo W, Zhang YH, Swietach P, et al. (2010) A mathematical model of the murine ventricular myocyte: a data-driven biophysically based approach applied to mice overexpressing the canine NCX isoform. Am J Physiol Heart Circ Physiol 299: H1045–63. [DOI] [PubMed] [Google Scholar]
  • 92. Mattiazzi A, Hove-Madsen L, Bers DM (1994) Protein kinase inhibitors reduce SR Ca transport in permeabilized cardiac myocytes. American Journal of Physiology - Heart and Circulatory Physiology 267: H812–H820. [DOI] [PubMed] [Google Scholar]
  • 93. Roth BJ (1997) Electrical conductivity values used with the bidomain model of cardiac tissue. Biomedical Engineering, IEEE Transactions on 44: 326–328. [DOI] [PubMed] [Google Scholar]
  • 94. Cordeiro JM, Spitzer KW, Giles WR, Ershler PE, Cannell MB, et al. (2001) Location of the initiation site of calcium transients and sparks in rabbit heart purkinje cells. The Journal of Physiology 531: 301–314. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95. Kitazawa T (1984) Effect of extracellular calcium on contractile activation in guinea-pig ventricular muscle. The Journal of Physiology 355: 635–659. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96. Backx PH, de Tombe PP, Van Deen JH, Mulder BJ, ter Keurs HE (1989) A model of propagating calcium-induced calcium release mediated by calcium diffusion. The Journal of General Physiology 93: 963–977. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97. Cascio W, Yan G, Kleber A (1992) Early changes in extracellular potassium in ischemic rabbit myocardium. the role of extracellular carbon dioxide accumulation and diffusion. Circulation Research 70: 409–422. [DOI] [PubMed] [Google Scholar]
  • 98. Swietach P, Rossini A, Spitzer KW, Vaughan-Jones RD (2007) H+ ion activation and inactivation of the ventricular gap junction: A basis for spatial regulation of intracellular pH. Circulation Research 100: 1045–1054. [DOI] [PubMed] [Google Scholar]
  • 99. Rodriguez B, Tice BM, Eason JC, Aguel F, Ferrero J Jose M, et al. (2004) Effect of acute global ischemia on the upper limit of vulnerability: a simulation study. Am J Physiol Heart Circ Physiol 286: H2078–2088. [DOI] [PubMed] [Google Scholar]
  • 100. Dennis SC, Gevers W, Opie LH (1991) Protons in ischemia: Where do they come from; where do they go to? Journal of Molecular and Cellular Cardiology 23: 1077–1086. [DOI] [PubMed] [Google Scholar]
  • 101. Bonen A (2001) The expression of lactate transporters (MCT1 and MCT4) in heart and muscle. European Journal of Applied Physiology 86: 6–11. [DOI] [PubMed] [Google Scholar]
  • 102. Vinnakota KC, Beard DA (2011) Kinetic analysis and design of experiments to identify the catalytic mechanism of the monocarboxylate transporter isoforms 4 and 1. Biophysical Journal 100: 369–380. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 103. Wang YP, Fuchs F (1994) Length, force, and Ca2+-troponin-C affinity in cardiac and slow skeletal-muscle. American Journal of Physiology 266: C1077–C1082. [DOI] [PubMed] [Google Scholar]
  • 104. Hak JB, van Beek JH, van Wijhe MH, Westerhof N (1993) Dynamics of myocardial lactate efflux after a step in heart rate in isolated rabbit hearts. American Journal of Physiology - Heart and Circulatory Physiology 265: H2081–H2085. [DOI] [PubMed] [Google Scholar]
  • 105. Rumsey WL, Pawlowski M, Lejavardi N, Wilson DF (1994) Oxygen pressure distribution in the heart in vivo and evaluation of the ischemic “border zone”. Am J Physiol Heart Circ Physiol 266: H1676–1680. [DOI] [PubMed] [Google Scholar]
  • 106. Khuri SF, Flaherty JT, O'Riordan JB, Pitt B, Brawley RK, et al. (1975) Changes in intramyocardial st segment voltage and gas tensions with regional myocardial ischemia in the dog. Circulation Research 37: 455–63. [DOI] [PubMed] [Google Scholar]
  • 107. Hillis LD, Khuri SF, Braunwald E, Kloner RA, Tow D, et al. (1979) Assessment of the efficacy of interventions to limit ischemic injury by direct measurement of intramural carbon dioxide tension after coronary artery occlusion in the dog. The Journal of Clinical Investigation 63: 99–107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108. Park CO, Xiao XH, Allen DG (1999) Changes in intracellular Na+ and pH in rat heart during ischemia: role of Na+/H+ exchanger. Am J Physiol Heart Circ Physiol 276: H1581–1590. [DOI] [PubMed] [Google Scholar]
  • 109. Elliott AC, Smith GL, Eisner DA, Allen DG (1992) Metabolic changes during ischaemia and their role in contractile failure in isolated ferret hearts. The Journal of Physiology 454: 467–490. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110. Wilde AA, Escande D, Schumacher CA, Thuringer D, Mestre M, et al. (1990) Potassium accumulation in the globally ischemic mammalian heart. a role for the atp-sensitive potassium channel. Circulation Research 67: 835–43. [DOI] [PubMed] [Google Scholar]
  • 111. Pike MM, Luo CS, Clark MD, Kirk KA, Kitakaze M, et al. (1993) NMR measurements of Na+ and cellular energy in ischemic rat heart: role of Na+-H+ exchange. Am J Physiol Heart Circ Physiol 265: H2017–2026. [DOI] [PubMed] [Google Scholar]
  • 112. Ten Hove M, Nederhoff MGJ, Van Echteld CJA (2005) Relative contributions of Na+/H+ exchange and Na+/HCO3 cotransport to ischemic Nai + overload in isolated rat hearts. Am J Physiol Heart Circ Physiol 288: H287–292. [DOI] [PubMed] [Google Scholar]
  • 113. van Echteld CJA, Kirkels JH, Eijgelshoven MHJ, van der Meer P, Ruigrok TJC (1991) Intracellular sodium during ischemia and calcium-free perfusion: A 23Na NMR study. Journal of Molecular and Cellular Cardiology 23: 297–307. [DOI] [PubMed] [Google Scholar]
  • 114. Imahashi K, Kusuoka H, Hashimoto K, Yoshioka J, Yamaguchi H, et al. (1999) Intracellular sodium accumulation during ischemia as the substrate for reperfusion injury. Circulation Research 84: 1401–1406. [DOI] [PubMed] [Google Scholar]
  • 115. Van Emous JG, Vleggeert-Lankamp CLAM, Nederhoff MGJ, Ruigrok TJC, Van Echteld CJA (2001) Postischemic Na+-K+-ATPase reactivation is delayed in the absence of glycolytic ATP in isolated rat hearts. Am J Physiol Heart Circ Physiol 280: H2189–2195. [DOI] [PubMed] [Google Scholar]
  • 116. Xiao XH, Allen DG (1999) Role of Na+/H+ exchanger during ischemia and preconditioning in the isolated rat heart. Circulation Research 85: 723–730. [DOI] [PubMed] [Google Scholar]
  • 117. El Banani H, Bernard M, Baetz D, Cabanes E, Cozzone P, et al. (2000) Changes in intracellular sodium and ph during ischaemia–reperfusion are attenuated by trimetazidine. Cardiovascular Research 47: 688–696. [DOI] [PubMed] [Google Scholar]
  • 118. Imahashi K, Nishimura T, Yoshioka J, Kusuoka H (2001) Role of intracellular Na+ kinetics in preconditioned rat heart. Circulation Research 88: 1176–1182. [DOI] [PubMed] [Google Scholar]
  • 119. Baetz D, Bernard M, Pinet C, Tamareille S, Chattou S, et al. (2003) Different pathways for sodium entry in cardiac cells during ischemia and early reperfusion. Molecular and Cellular Biochemistry 242: 115–120. [PubMed] [Google Scholar]
  • 120. Kleber A (1983) Resting membrane potential, extracellular potassium activity, and intracellular sodium activity during acute global ischemia in isolated perfused guinea pig hearts. Circulation Research 52: 442–450. [DOI] [PubMed] [Google Scholar]
  • 121. Hartmann M, Decking UKM (1999) Blocking Na+-H+ exchange by cariporide reduces Na+-overload in ischemia and is cardioprotective. Journal of Molecular and Cellular Cardiology 31: 1985–1995. [DOI] [PubMed] [Google Scholar]
  • 122. Yan GX, Chen J, Yamada KA, Klber AG, Corr PB (1996) Contribution of shrinkage of extracellular space to extracellular K+ accumulation in myocardial ischaemia of the rabbit. The Journal of Physiology 490: 215–228. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 123. Vanheel B, Hemptinne Ad (1992) Influence of KATP channel modulation on net potassium efflux from ischaemic mammalian cardiac tissue. Cardiovascular Research 26: 1030–1039. [DOI] [PubMed] [Google Scholar]
  • 124. Hill J, Gettes L (1980) Effect of acute coronary artery occlusion on local myocardial extracellular K+ activity in swine. Circulation 61: 768–778. [DOI] [PubMed] [Google Scholar]
  • 125. Kanda A, Watanabe I, Williams ML, Engle CL, Li S, et al. (1997) Unanticipated lessening of the rise in extracellular potassium during ischemia by pinacidil. Circulation 95: 1937–1944. [DOI] [PubMed] [Google Scholar]
  • 126. Mitani A, Kinoshita K, Fukamachi K, Sakamoto M, Kurisu K, et al. (1991) Effects of glibenclamide and nicorandil on cardiac function during ischemia and reperfusion in isolated perfused rat hearts. Am J Physiol Heart Circ Physiol 261: H1864–1871. [DOI] [PubMed] [Google Scholar]
  • 127. Wiegand V, Guggi M, Meesmann W, Kessler M, Greitschus F (1979) Extracellular potassium activity changes in the canine myocardium after acute coronary occlusion and the influence of beta-blockade. Cardiovascular Research 13: 297–302. [DOI] [PubMed] [Google Scholar]
  • 128. Yan GX, Joshi A, Guo D, Hlaing T, Martin J, et al. (2004) Phase 2 reentry as a trigger to initiate ventricular fibrillation during early acute myocardial ischemia. Circulation 110: 1036–1041. [DOI] [PubMed] [Google Scholar]
  • 129. Kimura J, Miyamae S, Noma A (1987) Identification of sodium calcium exchange current in single ventricular cells of guinea-pig. Journal of Physiology-London 384: 199–222. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 130. Kimura S, Bassett AL, Kohya T, Kozlovskis PL, Myerburg RJ (1986) Simultaneous recording of action potentials from endocardium and epicardium during ischemia in the isolated cat ventricle: relation of temporal electrophysiologic heterogeneities to arrhythmias. Circulation 74: 401–409. [DOI] [PubMed] [Google Scholar]
  • 131. Gasser RN, Vaughan-Jones RD (1990) Mechanism of potassium efflux and action potential shortening during ischaemia in isolated mammalian cardiac muscle. The Journal of Physiology 431: 713–741. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 132. Cole WC, McPherson CD, Sontag D (1991) ATP-regulated K+ channels protect the myocardium against ischemia/reperfusion damage. Circulation Research 69: 571–81. [DOI] [PubMed] [Google Scholar]
  • 133. Saito T, Sato T, Miki T, Seino S, Nakaya H (2005) Role of ATP-sensitive K+ channels in electrophysiological alterations during myocardial ischemia: a study using kir6.2-null mice. American Journal of Physiology - Heart and Circulatory Physiology 288: H352–H357. [DOI] [PubMed] [Google Scholar]
  • 134. Rouet R, Picard S, Libersa C, Ghadanfar M, Alabaster C, et al. (2000) Electrophysiological effects of dofetilide in an in vitro model of “border zone” between normal and ischemic/reperfused myocardium. Circulation 101: 86–93. [DOI] [PubMed] [Google Scholar]
  • 135. Yvon A, Hanouz J, Terrien X, Ducouret P, Rouet R, et al. (2002) Electrophysiological effects of morphine in an in vitro model of the ‘border zone’ between normal and ischaemic–reperfused guinea-pig myocardium. British Journal of Anaesthesia 89: 888–895. [DOI] [PubMed] [Google Scholar]
  • 136. Ducroq J, Rouet R, Puddu PE, Sall L, Tabourel C, et al. (2005) Electrophysiological effects of azimilide in an in vitro model of simulated-ischemia and reperfusion in guinea-pig ventricular myocardium. European Journal of Pharmacology 518: 165–174. [DOI] [PubMed] [Google Scholar]
  • 137. Wikswo Jr JP, Lin SF, Abbas RA (1995) Virtual electrodes in cardiac tissue: a common mechanism for anodal and cathodal stimulation. Biophysical Journal 69: 2195–2210. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 138. Muzikant AL, Henriquez CS (1998) Validation of three-dimensional conduction models using experimental mapping: are we getting closer? Progress in Biophysics & Molecular Biology 69: 205–223. [DOI] [PubMed] [Google Scholar]
  • 139.Richardson G (2009) A multiscale approach to modelling electrochemical processes occurring across the cell membrane with application to transmission of action potentials. Mathematical Medicine and Biology. [DOI] [PubMed]
  • 140. Mori Y, Fishman GI, Peskin CS (2008) Ephaptic conduction in a cardiac strand model with 3D electrodiffusion. Proceedings of the National Academy of Sciences 105: 6463–6468. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 141. Hand P, Peskin C (2010) Homogenization of an electrophysiological model for a strand of cardiac myocytes with gap-junctional and electric-field coupling. Bulletin of Mathematical Biology 72: 1408–1424. [DOI] [PubMed] [Google Scholar]
  • 142. Nygren A, Halter JA (1999) A general approach to modeling conduction and concentration dynamics in excitable cells of concentric cylindrical geometry. Journal of Theoretical Biology 199: 329–358. [DOI] [PubMed] [Google Scholar]
  • 143. Vigmond EJ, Aguel F, Trayanova NA (2002) Computational techniques for solving the bidomain equations in three dimensions. Biomedical Engineering, IEEE Transactions on 49: 1260–1269. [DOI] [PubMed] [Google Scholar]
  • 144. Blanchard E, Solaro R (1984) Inhibition of the activation and troponin calcium binding of dog cardiac myofibrils by acidic ph. Circulation Research 55: 382–391. [DOI] [PubMed] [Google Scholar]
  • 145. Orchard CH, Kentish JC (1990) Effects of changes of pH on the contractile function of cardiac-muscle. American Journal of Physiology 258: C967–C981. [DOI] [PubMed] [Google Scholar]
  • 146. Mitani A, Shattock MJ (1992) Role of Na-activated K channel, Na-K-Cl cotransport, and Na-K pump in [K]e changes during ischemia in rat heart. Am J Physiol Heart Circ Physiol 263: H333–340. [DOI] [PubMed] [Google Scholar]
  • 147. Coronel R, Wilms-Schopman FJ, Opthof T, van Capelle FJ, Janse MJ (1991) Injury current and gradients of diastolic stimulation threshold, TQ potential, and extracellular potassium concentration during acute regional ischemia in the isolated perfused pig heart. Circulation Research 68: 1241–9. [DOI] [PubMed] [Google Scholar]
  • 148. Hearse DJ, Yellon DM (1981) The border zone in evolving myocardial infarction: Controversy or confusion? The American Journal of Cardiology 47: 1321–1334. [DOI] [PubMed] [Google Scholar]
  • 149. Schaapherder AFM, Schumacher CA, Coronel R, Fiolet JWT (1990) Transmural inhomogeneity of extracellular K. and pH and myocardial energy metabolism in the isolated rat heart during acute global ischemia; dependence on gaseous environment. Basic Research in Cardiology 85: 33–44. [DOI] [PubMed] [Google Scholar]
  • 150. Prinzen FW, Arts T, Hoeks APG, Reneman RS (1989) Discrepancies between myocardial blood flow and fiber shortening in the ischemic border zone as assessed with video mapping of epicardial deformation. Pflügers Archiv European Journal of Physiology 415: 220–229. [DOI] [PubMed] [Google Scholar]
  • 151. Carmeliet E (2006) Action potential duration, rate of stimulation, and intracellular sodium. Journal of Cardiovascular Electrophysiology 17: S2–S7. [DOI] [PubMed] [Google Scholar]
  • 152. Wilde AA, Aksnes G (1995) Myocardial potassium loss and cell depolarisation in ischaemia and hypoxia. Cardiovascular Research 29: 1–15. [PubMed] [Google Scholar]
  • 153. Kleber AG, Janse MJ, van Capelle FJ, Durrer D (1978) Mechanism and time course of S-T and T-Q segment changes during acute regional myocardial ischemia in the pig heart determined by extracellular and intracellular recordings. Circulation Research 42: 603–613. [DOI] [PubMed] [Google Scholar]
  • 154. Kantor PF, Coetzee WA, Carmeliet EE, Dennis SC, Opie LH (1990) Reduction of ischemic K+ loss and arrhythmias in rat hearts. effect of glibenclamide, a sulfonylurea. Circulation Research 66: 478–85. [DOI] [PubMed] [Google Scholar]
  • 155. Robertson SP, Johnson JD, Potter JD (1981) The time-course of Ca2+ exchange with calmodulin, troponin, parvalbumin, and myosin in response to transient increases in Ca2+ . Biophysical Journal 34: 559–569. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 156. Allen DG, Orchard CH (1983) The effects of changes of pH on intracellular calcium transients in mammalian cardiac-muscle. Journal of Physiology 335: 555–567. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 157. Vaughan-Jones RD, Spitzer KW, Swietach P (2009) Intracellular pH regulation in heart. Journal of Molecular and Cellular Cardiology 46: 318–331. [DOI] [PubMed] [Google Scholar]
  • 158. Bollensdorff C, Knopp A, Biskup C, Zimmer T, Benndorf K (2004) Na+ current through katp channels: consequences for Na+ and K+ fluxes during early myocardial ischemia. Am J Physiol Heart Circ Physiol 286: H283–295. [DOI] [PubMed] [Google Scholar]
  • 159. Eigel BN, Hadley RW (1999) Contribution of the Na+ channel and Na+/H+ exchanger to the anoxic rise of Na+. in ventricular myocytes. American Journal of Physiology 277: H1817–1822. [DOI] [PubMed] [Google Scholar]
  • 160. Ellis D (1977) Effects of external cations and ouabain on intracellular sodium activity of sheep heart purkinje-fibers. Journal of Physiology-London 273: 211–240. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 161. Deitmer JW, Ellis D (1980) Interactions between the regulation of the intracellular pH and sodium activity of sheep cardiac purkinje fibres. The Journal of Physiology 304: 471–488. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 162. Bers DM, Ellis D (1982) Intracellular calcium and sodium activity in sheep heart purkinje-fibers - effect of changes of external sodium and intracellular pH. Pfluegers Archiv European Journal of Physiology 393: 171–178. [DOI] [PubMed] [Google Scholar]
  • 163. Katzung B (1975) Effects of extracellular calcium and sodium on depolarization-induced automaticity in guinea pig papillary muscle. Circulation Research 37: 118–127. [DOI] [PubMed] [Google Scholar]
  • 164. Baartscheer A, Schumacher CA, Fiolet JWT (1997) Small changes of cytosolic sodium in rat ventricular myocytes measured with SBFI in emission ratio mode. Journal of Molecular and Cellular Cardiology 29: 3375–3383. [DOI] [PubMed] [Google Scholar]
  • 165. Vaughan-Jones RD, Wu ML (1990) Extracellular H+ inactivation of Na+-H+ exchange in the sheep cardiac purkinje fibre. Journal of Physiology 428: 441–466. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 166. Ellis D, MacLeod KT (1985) Sodium-dependent control of intracellular pH in purkinje fibres of sheep heart. The Journal of Physiology 359: 81–105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 167. Kusuoka H, de Hurtado MCC, Marban E (1993) Role of sodium/calcium exchange in the mechanism of myocardial stunning: Protective effect of reperfusion with high sodium solution. Journal of the American College of Cardiology 21: 240–248. [DOI] [PubMed] [Google Scholar]
  • 168. Sasyniuk BI, Jhamandas V (1984) Mechanism of reversal of toxic effects of amitriptyline on cardiac purkinje fibers by sodium bicarbonate. Journal of Pharmacology and Experimental Therapeutics 231: 387–394. [PubMed] [Google Scholar]
  • 169. Cheung JY, Thompson IG, Bonventre JV (1982) Effects of extracellular calcium removal and anoxia on isolated rat myocytes. American Journal of Physiology - Cell Physiology 243: C184–C190. [DOI] [PubMed] [Google Scholar]
  • 170. Kaila K, Vaughan-Jones RD (1987) Influence of sodium-hydrogen exchange on intracellular pH, sodium and tension in sheep cardiac purkinje fibres. The Journal of Physiology 390: 93–118. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 171. Deitmer JW, Ellis D (1978) Changes in the intracellular sodium activity of sheep heart purkinje fibres produced by calcium and other divalent cations. The Journal of Physiology 277: 437–453. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 172. Allen DG, Eisner DA, Orchard CH (1984) Factors influencing free intracellular calcium concentration in quiescent ferret ventricular muscle. The Journal of Physiology 350: 615–630. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 173. Sheu S, Korth M, Lathrop D, Fozzard H (1980) Intra- and extracellular K+ and Na+ activities and resting membrane potential in sheep cardiac purkinje strands. Circulation Research 47: 692–700. [DOI] [PubMed] [Google Scholar]
  • 174. Spitzer KW, Walker JL (1980) Intracellular chloride activity in quiescent cat papillary muscle. American Journal of Physiology - Heart and Circulatory Physiology 238: H487–H493. [DOI] [PubMed] [Google Scholar]
  • 175. Deitmer JW, Ellis D (1980) The intracellular sodium activity of sheep heart purkinje fibres: effects of local anaesthetics and tetrodotoxin. The Journal of Physiology 300: 269–282. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 176. Miura DS, Hoffman BF, Rosen MR (1977) The effect of extracellular potassium on the intracellular potassium ion activity and transmembrane potentials of beating canine cardiac purkinje fibers. The Journal of General Physiology 69: 463–474. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 177. Abete P, Vassalle M (1992) Role of intracellular Na+ activity in the negative inotropy of strophanthidin in cardiac purkinje fibers. European Journal of Pharmacology 211: 399–409. [DOI] [PubMed] [Google Scholar]
  • 178. Vaughan-Jones RD (1979) Regulation of chloride in quiescent sheep-heart purkinje fibres studied using intracellular chloride and pH-sensitive micro-electrodes. J Physiol (Lond) 295: 111–137. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 179. Cyran SE, Phillips J, Ditty S, Baylen BG, Cheung J, et al. (1993) Developmental differences in cardiac myocyte calcium homeostasis after steady-state potassium depolarization: Mechanisms and implications for cardioplegia. The Journal of Pediatrics 122: S77–S83. [DOI] [PubMed] [Google Scholar]
  • 180. Chapman RA (1986) Sodium/calcium exchange and intracellular calcium buffering in ferret myocardium: an ion-sensitive micro-electrode study. The Journal of Physiology 373: 163–179. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 181. Vaughan-Jones RD (1986) An investigation of chloride-bicarbonate exchange in the sheep cardiac purkinje fibre. Journal of Physiology 379: 377–406. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 182. Bountra C, Powell T, Vaughan-Jones RD (1990) Comparison of intracellular pH transients in single ventricular myocytes and isolated ventricular muscle of guinea-pig. The Journal of Physiology 424: 343–365. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 183. Gambassi G, Hansford RG, Sollott SJ, Hogue BA, Lakatta EG, et al. (1993) Effects of acidosis on resting cytosolic and mitochondrial Ca2+ in mammalian myocardium. The Journal of General Physiology 102: 575–597. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 184. Fozzard HA, Lee CO (1976) Influence of changes in external potassium and chloride ions on membrane potential and intracellular potassium ion activity in rabbit ventricular muscle. J Physiol 256: 663–689. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 185. Carmeliet E, Verdonck F (1977) Reduction of potassium permeability by chloride substitution in cardiac cells. The Journal of Physiology 265: 193–206. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 186. Wu M, Tsai M, Tseng Y (1994) DIDS-sensitive pHi regulation in single rat cardiac myocytes in nominally HCO3 free conditions. Circulation Research 75: 123–132. [DOI] [PubMed] [Google Scholar]
  • 187. Bolton TB, Vaughan-Jones RD (1977) Continuous direct measurement of intracellular chloride and ph in frog skeletal muscle. The Journal of Physiology 270: 801–833. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 188. Bountra C, Vaughan-Jones RD (1989) Effect of intracellular and extracellular pH on contraction in isolated, mammalian cardiac muscle. The Journal of Physiology 418: 163–187. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 189. Miura D, Rosen M (1978) The effects of ouabain on the transmembrane potentials and intracellular potassium activity of canine cardiac purkinje fibers. Circulation Research 42: 333–338. [DOI] [PubMed] [Google Scholar]

Articles from PLoS ONE are provided here courtesy of PLOS

RESOURCES