Noble2002-Modelling the Heart-Insights Failures and Progress

download Noble2002-Modelling the Heart-Insights Failures and Progress

of 9

Transcript of Noble2002-Modelling the Heart-Insights Failures and Progress

  • 8/8/2019 Noble2002-Modelling the Heart-Insights Failures and Progress

    1/9

    Modelling the heart: insights,failures and progressDenis Noble

    SummaryMathematical models of the heart have developed overa period of about 40 years. Cell types in all regions ofthe heart have been modelled and they are now beingincorporated into anatomically detailed models of thewhole organ. This combination is leading to the creationof the first virtual organ, which is being used in drugdiscovery and testing, and in simulating the action ofdevices, such as cardiac defibrillators. Simulation is anecessary tool of analysis in attempting to understandbiological complexity. We often learn as much from thefailures as from the successes of mathematical models.It is the iterative interaction between experiment and

    simulation that is important. Examples are given wherethis process has been instrumental in some of the majoradvances in the field. BioEssays24:11551163, 2002. 2002 Wiley Periodicals, Inc.

    Introduction

    Mathematical modelling is widely accepted as an essential

    tool of analysis in the physical sciences and engineering, yet

    many arestill scepticalaboutits role in biology. Why is this so?

    A partial answer would be that many chose to work as biolo-

    gistsprecisely because theywere not mathematically inclined.

    But I think there is a stronger reason. There is also a wide-

    spread belief in the biological sciences that, to be useful,

    theory must be correct. Failure is seen as just that, i.e. failure,

    rather than as an essential part of a two-way iterative process

    between theory and experiment.

    In the case of excitable cells, it could also be that the para-

    digm model, the Hodgkin Huxley(1) equations for the squid

    nerve action potential, was so spectacularly successful that it

    created an unrealistic expectation for the rapid application of

    the same principles elsewhere. In fact, modelling of the much

    more complex cardiac cell has required many years of inter-

    action between experiment and theory. In modelling com-

    plex biological phenomena, this is precisely what we should

    expect(2,3) and it is standard for such interaction to occur over

    many years in thephysicalsciencesand engineering.I will first

    illustrate this interaction in biological simulation using some

    of the cell models that I have been involved in developing.

    Since my purpose is didactic, I will be highly selective. A more

    complete historical review of cardiac cell models can be found

    elsewhere(4) and the volume in which that article appeared is

    also a rich sourceof material on modellingthe heart,since that

    was its focus.

    At each stage, I will highlight the theoretical insights that

    came from mathematical modelling as well as the experi-

    mental basis. I will then briefly review the use of cell models in

    tissue and organ simulations.

    Energy conservation during the cardiac cycle:

    discovery of various forms of potassium

    channels

    The first cardiac cell models sought insight into the most

    obvious difference between electrical activity in heart and

    nerve: the duration of the action potential. A nerve action

    potential may last only 1 msec. Its function is to encode infor-

    mation as rapidly as possible. In contrast, a human ventricular

    action potential may last 400 msec, during which time many

    events are triggered that initiate and control mechanical

    contraction.

    FitzHugh(5)

    showed that the HodgkinHuxley model ofthe nerve impulse could generate slow repolarization by

    greatly reducing the amplitude and speed of activation of the

    delayed potassiumcurrent,IK. These changes notonlyslowed

    repolarization; they also created a plateau. This result showed

    that there must be some property inherent in the Hodgkin

    Huxley formulation of the sodium current that permits a

    persistent inward current to occur. The main defect of the

    FitzHugh model was that it was a very expensive way of

    generating a plateau, with such high ionic conductances that,

    during each action potential, the sodium and potassium ionic

    gradients would be run down at a rate at least an order of

    magnitude too large.

    That this was not the case was already evident sinceWeidmanns(6,7) pioneering work showed that the conduc-

    tance during the action potential is very low. The experimental

    reason for this became clear with the discovery of the inward-

    rectifier current, IK1(810) (see Fig. 1, top). The permeability of

    this channel falls almost to zero during strong depolarisation.

    These experiments were also thefirst to show that there are at

    least two K conductances in the heart, IK1 and IK (referred to

    BioEssays 24:11551163, 2002 Wiley Periodicals, Inc. BioEssays 24.12 1155

    University Laboratory of Physiology, Parks Road, Oxford OX1 3PT,

    UK. E-mail: [email protected]

    Funding agencies: The British Heart Foundation, The Medical

    Research Council, The Wellcome Trust and Physiome Sciences Inc.

    DOI 10.1002/bies.10186

    Published online in Wiley InterScience (www.interscience.wiley.com).

    Review articles

  • 8/8/2019 Noble2002-Modelling the Heart-Insights Failures and Progress

    2/9

    as IK2 in early work, but now known to consist of IKr and IKs).

    The Noble 1962 model(11,12) (Fig. 1, bottom) was construct-

    ed to determine whether this combination of K channels,

    togetherwith a Hodgkin Huxleytype(see Box 1 in Marder and

    Prinz in this issue of BioEssays) sodiumchannel could explain

    all the classical Weidmann experiments on conductance

    changes. The model not only succeeded in doing this; it also

    demonstrated that an energy-conserving plateau mechanism

    was an automatic consequence of the inward-rectifying pro-

    perties of IK1. This has featured in all subsequent models,

    and it is a very important insight. The main advantage of a low

    conductance is minimising energy expenditure.

    Unfortunately, however, a low conductance plateau was

    achieved at the cost of making the repolarization process

    fragile.Pharmaceutical companies today are struggling to deal

    with evolutions answer to this problem, which was to entrust

    repolarization to the potassium channel iKr. This channel pro-

    tein is one of the most promiscuous receptors known: large

    ranges of drugs can enter the channel mouth to block it, and

    even more interact with the G-protein-coupled receptors thatcontrol it. The consequence can be failed repolarization, and

    the triggering of fatal arrhythmias (see http://georgetowncert.

    org/qtdrugs_torsades.asp). Computer simulation is now play-

    ing a major role in attempting to find a way around this difficult

    and seemingly intractable problem.(13)

    The main defect of the 1962 model was that it included

    only one voltage-gated inward current, INa. There was a good

    reason for this. Calcium currents had not then been dis-

    covered. There was, nevertheless, a clue in the model that

    something important was missing. The only way in which

    the model could be made to work was to greatly extend the

    voltage range of the sodium window current by reducing

    thevoltage dependence of thesodium activation process (seeRef. 12, Figure 15). In effect, the sodium current was made to

    serve thefunctionof both thesodium and calcium channels so

    far as the plateau is concerned. There was a clear prediction

    here: either sodium channels in the heart are quantitatively

    different from those in nerve, or other inward current-carrying

    channels must exist. Both predictions are correct.

    The first successful voltage clamp measurements came in

    1964(14) and they rapidly led to the discovery of the cardiac

    calcium current.(15) By the end of the 1960s therefore, it was

    already clear that the 1962 model needed replacing.

    Controversy over the pacemaker

    current: the MNT modelIn addition to the discovery of the calcium current, the early

    voltage clamp experiments also revealed multiple compo-

    nents of IK. Noble & Tsien(16) labelledthem Ix1 and Ix2, but they

    are now referred to as IKr and IKs.(17) They also showed that

    these slow gated currents in the plateau range of potentials

    were quite distinct from those near the resting potential, i.e.

    that there were twoseparate voltage rangesin which very slow

    Figure 1. Top: Experimental basis of the first classification

    of potassium currents. Redrawn from reference (10, Fig. 2B).

    The solid line shows the total membrane current recorded in

    a Purkinje fibre in a sodium-depleted solution. The inward-

    rectifying current was identified as iK1, which is extrapolated

    here as nearly zero at positive potentials. The outward-

    rectifying current, IK, is now known to be mostly formed by

    the component IKr. Thehorizontal arrow indicatesthe trajectory

    at the beginning of the action potential, while the vertical

    arrow indicates the time-dependent activation of IK which

    initiates repolarization. Bottom: Sodium and potassium

    conductance changes computed from the 1962 model(12) of

    the Purkinje fibre. Two cycles of activity are shown. The con-

    ductances are plotted on a logarithmic scale to accommodate

    the large changes in sodium conductance. Note the persistentlevel of sodium conductance during the plateau of the action

    potential, which is about 2% of the peak conductance. Note

    also therapid fall in potassium conductanceat thebeginningof

    the action potential. This is attributable to the properties of the

    inward rectifier iK1.

    Review articles

    1156 BioEssays 24.12

  • 8/8/2019 Noble2002-Modelling the Heart-Insights Failures and Progress

    3/9

    conductance changes could be observed.(16,18) These experi-

    ments formed the basis of the MNT model.(19)

    The MNT model reconstructed a much wider range of

    experimentalresults, andit didso with great accuracy in some

    cases. A good example of this was the reconstruction of the

    paradoxical effect of small current pulses on the pacemaker

    depolarisation in Purkinje fibres (see Fig. 2)paradoxical

    because brief depolarisations slow the process and brief

    hyperpolarizations greatly accelerate it. Reconstructing para-

    doxical or counter-intuitive results is of coursea major function

    of modelling work. This is one of the roles of modelling in

    unravelling complexity in biological systems.

    But the MNT model also contained the seeds of a spec-

    tacular failure. Following the experimental evidence,(18) it

    attributed the slow conductance changes near the resting

    potential to a slow gated potassium current, IK2. In fact, what

    became the pacemaker current, or If, is an inward current

    activated by hyperpolarization(20) not an outward current acti-

    vated by depolarisation. At thetime, it seemed hard to imagine

    a more serious failure than getting both the current directionand the gating by voltage completely wrong. There cannot be

    much doubt therefore that this stage in the iterative interaction

    between experiment and simulation created a major problem

    of credibility. Perhaps cardiac electrophysiology was not really

    ready for modelling work to be successful?

    This was how the failure was widely perceived. Yet that

    perception was a deep misunderstanding of the significance

    of what was emerging from this experience. It was no

    coincidence that both the current direction and the gating

    were wrong, as one follows from the other. And so did much

    else in themodelling!Working that outin detailwas theground

    on which future progress could be made.

    This is thepointat which to emphasise one of theimportant

    points about the philosophy of modelling which I highlighted

    at the beginning of this article: it is one of the functions of

    models to be wrong! (Indeed, a Popperian philosopher of

    science would say that one can only ever establish the falsity

    of theories.) Of course, there are many ways of being wrong,

    and I am not talking here of failing in arbitrary or purely con-

    tingent ways, but in ways that advance our understanding by

    exploring possible logics of complex systems and determining

    which are most accurate. Again, this situation is familiar to

    those working in simulation studies in engineering or cosmol-

    ogy or in many other physical sciences.And, in fact, thefailure

    of the MNT model is one of the most instructive examples

    of experiment simulation interaction in physiology, and of

    subsequent successful model development. I do not have the

    space here to review this issue in all its details. That has

    already been done.(21,22) Here I will simply draw the conclu-

    sions relevant to modern work.

    First, careful analysis of the MNT model revealed that its

    pacemaker current mechanism could not be consistent withwhat isknownof theprocessof ionaccumulationanddepletion

    in the extracellular spaces between cells. The model itself

    was therefore a key tool in understanding the next stage of

    development.

    Second, a complete and accurate mapping between the

    IK2 model and the new If model could be constructed(21)

    (see Fig. 2) demonstrating how both models related to the

    same experimental results and to each other. Such mapping

    between different modelsis rare inbiologicalwork,but itcan be

    very instructive.

    Third, this spectacular turn-around was the trigger for

    the development of models that include changes in ion

    Figure 2. Reconstructionof theparadoxical effect ofsmall currentsinjectedduringpacemakeractivity.Left: Computations fromthe MNT

    model.(19) Small depolarising and hyperpolarizing currents were applied for 100 msec during the middle of the pacemaker depolarisation.

    Hyperpolarizations are followed by an acceleration of the pacemaker depolarisation, whilesubthreshold depolarisations induce a slowing.

    Middle: Experimental records from Weidmann (Ref. 6, Fig. 3). Right: Similar computations using the DiFrancesco-Noble (1985)

    model.(24) Despite thefundamental differencesbetweenthese twomodels,the featurethat explainsthe paradoxical effects of small current

    pulses survives. This kind of detailed comparison was part of the process of mapping the two models onto each other.

    Review articles

    BioEssays 24.12 1157

  • 8/8/2019 Noble2002-Modelling the Heart-Insights Failures and Progress

    4/9

    concentrations inside and outside the cell and between

    intracellular compartments.

    Finally, the MNT model was the point of departure for the

    ground-breaking work of Beeler and Reuter(23) who developed

    the first ventricular cell model. As they wrote of their model:

    in a sense, it forms a companion presentation to the recent

    publication of McAllister et al. (1975) on a numerical re-

    construction of the cardiac Purkinje fibre action potential.

    There are sufficiently many and important differences be-

    tween these two types of cardiac tissue, both functionally and

    experimentally, that a more or less complete picture of mem-

    brane ionic currents in the myocardium must include both

    simulations. For a recent assessment of this model and the

    subsequent Luo-Rudy models see Ref. 4.

    Ion concentrations, pumps and exchangers:

    The DiFrancescoNoble model(24)

    The incorporation not only of ion channels (following the

    HodgkinHuxley paradigm) but also of ion exchangers, such

    as NaK exchange (the sodium pump), NaCa exchange,the SR calcium pump and, more recently, the transporters

    involved in controlling cellular pH,(25) was a fundamental

    advance since these are essential to the study of some dis-

    ease statessuch as congestive heart failure(26) and ischaemic

    heart disease.

    It was necessary to incorporate the NaK exchange pump

    since what made If so closely resemble a potassium channel

    in Purkinje fibres was the depletion of K ions in extracel-

    lular spaces. This was a key feature enabling the accurate

    mapping of the IK2 model (MNT) onto the If model.(21) But, to

    incorporate changes in ion concentrations it became neces-

    sary to represent the processes by which ion gradients can be

    restored and maintained. In a form of modelling avalanche,once changes in one cation concentration gradient (K) had

    been introduced, the others (Na and Ca2) had also to be

    incorporatedsince thechangesare alllinked viathe NaK and

    Na Ca exchange mechanisms. This avalanche of additional

    processes was the basis of the DiFrancescoNoble (1985)

    Purkinje fibre model.(24)

    The greatly increased complexity of the DiFrancesco

    Noble model, which for the first time also represented intra-

    cellular events by incorporating a model of calcium release

    from the sarcoplasmic reticulum, increased both the range of

    predictions and the opportunities for failure. Here I will limit

    myself to one example of each.

    The most influential prediction was that relating to thesodiumcalcium exchanger. In the early 1980s, it was still

    widely thought that the original electrically neutral stoichio-

    metry (Na:Ca2:1) derived from early flux measurements

    was correct. The DiFrancesco-Noble model achieved two

    important conclusions. The first was that, with the experimen-

    tally known Na gradient, there simply wasnt enough energy

    in a neutral exchanger to keep resting intracellular calcium

    levels below 1 mM, i.e. at a level low enough to permit relax-

    ation to occur. Switching to a stoichiometry of 3:1 readily

    allowed resting calcium to be maintained below 100 nM. This

    automatically led to the prediction that there must be a current

    carried by the NaCa exchanger and that, if this exchanger

    was activated by intracellular calcium, it must also be strongly

    time-dependent since intracellular calcium varies by an order

    of magnitude during each action potential. Even as the model

    was being published, experiments demonstrating the current

    INaCa were being performed(27) andthe variation of this current

    during activity was being revealed either as a late compo-

    nent of inward current or as a current tail on repolarization.

    This prediction has turned out to have very important con-

    sequences for the elucidation of some of the mechanisms of

    cardiac arrhythmia in disease states in whichcells accumulate

    sodium and calcium, either through loss of energy supply, as

    in ischaemia, or as a consequence of reduced activity of the

    NaK ATPase (Na pump) as in treatment with cardiac

    glycosides. At a critical level of sodiumand calcium accumula-

    tion, calcium release occurs spontaneously and becomesrepetitive between sodium concentrations around 13 and

    22 mM,(2831) a phenomenon also seen experimentally.

    Each calcium release activates inward current carried by

    sodium calcium exchange which, if large enough, can trigger

    additional (ectopic) action potentials,(32) as shown in Fig. 3.

    The main defect of the DiFrancescoNoble model was

    that the intracellular calcium transient was far too large, mainly

    because the model did not represent the attachment of

    calcium to intracellular proteins. This signalled the need to

    incorporate intracellular calcium buffering.

    Calcium balance: the Hilgemann-Noble model

    This deficiency was tackled in the HilgemannNoble(33)

    (1987) modelling of the atrial action potential. Although this

    was directed towards atrial cells, it also provided a basis for

    modelling ventricular cells in species (rat, mouse) with short

    ventricular action potentials, and many of its features were

    adopted in later ventricular cell models of species with high

    plateaus.(3438)

    The HilgemannNoble model addressed a number of

    important questions concerning calcium balance:

    1. When does the calcium that enters during each action

    potential return to the extracellular space? Does it do this

    during diastole (as most people had presumed) or during

    systole itself, i.e. during, not after, the action potential?Hilgemann(39) had done experiments with tetramethylmur-

    exide, a calcium indicator restricted to the extracellular

    space, showing that the recovery of extracellular calcium

    (inintercellular clefts) occurs remarkably quickly (seeFig. 4,

    inset). In fact, net calcium efflux is established as soon as

    20 msec after thebeginning of theaction potential, which at

    that time was considered to be surprisingly soon. Calcium

    Review articles

    1158 BioEssays 24.12

  • 8/8/2019 Noble2002-Modelling the Heart-Insights Failures and Progress

    5/9

    activation of efflux via the Na Ca exchanger achieved this

    in the model (see Fig. 4compare the computed trace

    [Ca]0 with the experimental trace labelled [Ca]0)

    2. Where was the current that this would generate and did it

    correspond to the quantity of calcium that the exchanger

    needed to pump? Mitchell et al(40) had already done

    experiments in rat ventricle showing that replacement of

    sodium with lithium removes the late plateau. This was the

    first experimental evidence that the late plateau in action

    potentials with this shape might be maintained by sodium

    calcium exchange current. The HilgemannNoble model

    showed that this is precisely what one would expect.

    3. Could a model of the SR that reproduces at least the

    major features of Fabiatos experiments(41,42) showing

    calcium-induced calcium release (CICR) be incorporated

    into the cell models and integrate in with whatever were

    the answers to questions 1 and 2. This was a major

    challenge (see reference (33), pp 173174). The model

    followed as much of the Fabiato data as possible, but the

    conclusions were that the modelling, while broadly con-

    sistent with the Fabiato work, could not be based on that

    alone. It is an important function of simulation to reveal

    when experimental data needs extending.

    4. Were the quantities of calcium, free and bound, at each

    stage of the cycle consistent with the properties of the

    cytosol buffers? The answer here was a very satisfactory

    yes. The great majority of the cytosol calcium is bound so

    that, althoughmuch morecalciummovement was involved,

    the free calcium transients were much smaller, within theexperimental range.

    There were however some gross inadequacies in the

    calcium dynamics. An additional voltage-dependence of Ca

    release was inserted to obtain a fast calcium transient. This

    was a compromise that really requires proper modelling of

    the subsarcolemmal space where calcium channels and the

    Figure 3. Left: Calciumoscillations computedin a model of thePurkinje fibre during therise in intracellular sodium(red) followingblockof

    the sodium-potassium ATPase. Each oscillation of intracellular calcium (blue trace below) triggers inward sodium-calcium exchange

    current (brown trace top) (from Ref. 25). These computations were done under voltage clamp conditions. Right: Free-running action

    potentials in thesame model withintracellularsodiumset at a level corresponding to themiddleof theoscillationperiodin theleft handfigure

    (from Ref. 32). For a theoretical analysis of this phenomenon in the DiFrancescoNoble model see Refs. 28,29 and for an analysis of the

    more detailed WinslowRiceJafri model(30) see Ref. 31.

    Figure 4. The first reconstruction of calcium balance in

    cardiac cells. The HilgemannNoble model(33) incor-

    porated complete calcium cycling, such that intracellular and

    extracellular calcium levels returned to their original state after

    each cycleand that the effects of sudden changes in frequency

    could be reproduced. Left: Computed action potential (AP),

    intracellular calcium transient, contraction (represented by

    cross-bridge formation) and extracellular calcium transient.

    Inset: Experimental recording of action potential (AP), cell

    motion and extracellular calcium transient.

    Review articles

    BioEssays 24.12 1159

  • 8/8/2019 Noble2002-Modelling the Heart-Insights Failures and Progress

    6/9

    ryanodine receptors interact, a problem later tackled by Jafri,

    Rice & Winslow(43,44) and by Noble et al.(38) Another problem

    was how would the conclusions apply to action potentials with

    high plateaus. This was tackled both experimentally(45) and

    computationally.(35,38) The answer is that the high plateau in

    ventricular cells of guinea-pig, dog, human etc greatly delays

    the reversal of the sodium calcium exchanger so that net

    calcium entry continues for a longer fraction of the action

    potential. This property is important in determining the force-

    frequency characteristics.

    Incorporation of cell models into tissue

    and organ models

    Although single-cell models of the heart were first constructed

    over 40 years ago, it is only relatively recently that it has

    become possible to move up to the next levels: tissue and

    organ models into which the cell models can be incorporated.

    This was always one of the aims of the single cell work, but

    until the development of sufficiently powerful supercomputers

    this objective was beyond reach.One of the early examples of massive three-dimensional

    ionic tissue models usedintracellular calciumoscillator models

    of the kind shown in Fig. 3 to address the question of the

    minimal size of an ischaemic focus, simulated by sodium

    overload in a defined tissue region, that would be arrhyth-

    mogenic.(46) A sodium-overloaded cell group was placed in a

    20-cell-diameter spherical volume in the centre of a three-

    dimensional network of 12812832 atrial cells using the

    EarmNoble single cellmodel(47) (i.e.more than500,000ionic

    cell models were used). It was shown that this relatively small

    ischaemic focusin real terms, the tissue volume would be

    in the order of no more than a cubic millimetreis capable of

    triggering a spreading wave of excitation in the whole tissue

    model (Fig. 5). This was an important step forward in link-

    ing a subcellular mechanism of arrhythmia to multicellular

    processes. However, the initiation of an ectopic beat is not,

    in itself, sufficient to trigger a life-threatening arrhythmia

    (fortunately most ectopic beats are benign!) It still remains to

    work out how this mechanism interacts with other features of

    ischemic tissue, such as slowed conduction due to potassium

    accumulation (4850) and action potential shortening, to trigger

    re-entrant arrhythmias that may break down into full-scale

    ventricular fibrillation.

    On the one hand, this example shows that both qualitative

    andquantitativeconclusionsmay be drawnfrom computations

    based on multidimensional, uniform grid-structured cardiac

    models. On the other hand, the deficiencies in these models

    also show that geometrical factors, tissue (in)homogeneity,

    cell-to-cell coupling, etc., are critical to an understanding of

    detailedphysiology and pathology. Thus, while grid-structured

    cardiac tissue models clearly have their place, more sophis-

    ticated anatomical models are required to progress be-yond the reproduction of biophysical behaviour and towards

    the provision of physiologically and clinically relevant new

    insights.

    Simulating myocardial excitation in three-dimensional

    anatomically detailed models of the heart began with the work

    of Panfilov and colleagues(51,52) who mapped finite difference

    models of FitzHughNagumo equations to experimentally

    determined cardiac geometry and fibre distribution data.(53)

    The first solutions of ionic current models used finite

    difference techniques and non-deforming anatomical heart

    geometry.(54)Later, ionic current models were linked to mate-

    rial points of a moving finite element mesh.(55) This allowed

    Figure 5. A three-dimensional atrial model with

    centrally located small sodium-overloaded ischae-

    mic focus. The dimensions of the atrial tissue block

    are outlined in orange, while the action potential

    wavefront and plateau are colour coded (red and

    magenta,respectively).The excitation begins in the

    small central region of sodium overload, where a

    calcium oscillation generates an action potential by

    activating sodium-calcium exchange (see Fig. 3).

    If the overload region (top left) is sufficiently large,

    the current it generates is capable of exciting the

    surrounding normal tissue (top right), so initiating

    a propagated wave (bottom left). Reprinted from

    Winslow RL, Varghese A, Noble D, Adlakha C,

    Hoythya A. Proc Roy Soc B 1990;240:8396, with

    permission from The Royal Society.

    Review articles

    1160 BioEssays 24.12

  • 8/8/2019 Noble2002-Modelling the Heart-Insights Failures and Progress

    7/9

    simulation of the electromechanical coupling between the

    activation process and the mechanically deforming finite

    element model in both directions: the solution of the electro-

    physiological equations described the calcium current that

    initiates contraction, and the shape changes caused by

    cardiac contraction influenced electrical propagation.(56)

    Extensions of this work have nowbeenmadeto accommodate

    more-sophisticated models of cardiac ion channels(3638) and

    cellular electromechanical coupling.(57)

    At the same time, other important elements of a whole-

    heart model are being put in place. Modelling of the coronary

    circulation(58) has advanced to the point at which it is possible

    to simulate coronary occlusion and so define the areas in the

    ventricle that would become ischaemic.(32) Work is also in

    progress to incorporate the SA node, atrial and Purkinje

    conducting network. A reasonably complete virtual heart is

    well on the way to construction. Such a project, with immense

    implications for practical use in designing new therapies, (32)

    would have been inconceivable until very recently.

    I will finish this article with a spectacular example of thiskind of use of the virtual heart. Figure 6 (top) shows the

    anatomical model of the canine ventricles(59) witha 1 mmthick

    slice in which the fibre orientations are shown. The lower part

    of the figure shows three frames from a simulation of arrhy-

    thmia (called a tachycardia) in which a rotating spiral wave

    (at the back), triggered by the collision of two stimulated

    waves, is generating successive waves that propagate

    through the structure of the slice. This simulation uses a

    modified version of the BeelerReuter ventricular cell equa-

    tions.(60) It is part of a major study not only of the conditions for

    initiating arrhythmia but also of the mechanisms by which

    electrical defibrillation may work. The model is the first to use

    bi-domain simulation (in which the electrical field outsidethe cells is reconstructed as well as the potentials across

    the cell membranes), withdetailedfibre geometry and detailed

    cellular electrophysiology.(61,62)

    Such simulations require enormous computing resources,

    currently available only on large supercomputers (which is

    the main reason why the simulation is restricted to a slice of

    the ventriclecomputations of the whole canine ventricle

    would, at present, be too demanding).

    Modularity in biological systems

    A recurring theme in these developments is what I have

    called a modelling avalanche in which incorporation of one

    new element has required incorporation of a set of relatedprocesses. Biological modelling often exhibits this form of

    modularity, making it necessary to incorporate a group of

    protein components together. It will be one of the major chal-

    lenges of mathematical biology to use simulation work to

    unravel the modularity of nature. Groups of proteins co-

    operating to generate a function and therefore being selected

    together in the evolutionary process will be revealed by this

    approach. This piecemeal approach to reconstructing the

    logic of life, which is the strict meaning of the word phy-

    siology,(63) could also be theroutethroughwhich a systematic

    theoretical biology could eventually emerge.(64,65)

    Figure 6. Top: Diagram of the canine ventricular model(59)

    showing the fibre structure of the slice in which the computa-

    tions shown below were carried out. The slice is represented

    as a full bi-domain model, i.e. the electric current flow in

    the extracellular space is also computed. This enables such

    models to be used to investigate the effects of massive electric

    shocks of the kind used in defibrillation. Bottom: Rotating

    spiral wave generating successive propagated waves com-

    putedwithin theslice modelusing a modification ofthe Beeler

    Reuter ventricular cell ionic current equations (from Refs.

    61,62).

    Review articles

    BioEssays 24.12 1161

  • 8/8/2019 Noble2002-Modelling the Heart-Insights Failures and Progress

    8/9

    Future challenges and the nature

    of biological simulation

    Progress in this field was slow and spasmodic in the period

    between 1960 and 1990. This reflected the lack of experi-

    mental data (a problem that was overcome at the cellular level

    with the introduction of the patch clamp technique and its

    application to isolated cells), andthe lack of computing power.

    But since 1990 there has been an extraordinary explosion of

    modelling work on the heart.(66) There are multiple models of

    all the cell types, and I confidently predict that there will be

    many more to come. Why do we have so many? Couldnt

    we simply standardise the field and choose the best? To

    some extent, that is happening. None of the historical models

    described in this article is now used much in its original form.

    One of the major reasons for the multiplicity of models is

    that there will always be a compromise between complexity

    and computability. A good example here is the modelling of

    calcium dynamics. As we understand these dynamics in ever

    greater detail,(31) models become more accurate and they

    encompass more biological data, but they also becomecomputationally demanding. This was the motivation behind

    the simplified dyadic space model of Noble et al. (38) which

    achieves many of the required features of the initiation of

    calcium signalling with only a modest (10%) increase in

    computation time, an important consideration when import-

    ing such models into models of the whole heart. But no-one

    would use that model to study the fine properties of calcium

    dynamics at the subcellular level. That was not its purpose.

    There will probably therefore be no unique model that does

    everything at all levels. Any of the boxes at one level could be

    deepened in complexity at a lower level, or fused with other

    processes at a higher level. In any case, all models are only

    partial representations of reality. One of the first questions toaskof a modeltherefore is what questionsdoes it answerbest.

    It is through the iterative interaction between experiment and

    simulation that we will gain that understanding.

    It is however already clear that incorporation of cell models

    into tissue and organ models is capable of spectacular

    insights. The incorporation of cell models into anatomically

    detailed heart models, as shown in Fig. 6, hasbeen an exciting

    development. The goal of creating an organ model capable of

    spanning the whole spectrum of levels from genes(6771) to

    the electrocardiogram(13,32) is within sight, and is one of the

    challenges of the immediate future. The potential of such

    simulations for teaching, drug discovery, device development

    and, of course, for pure physiological insight is only beginningto be appreciated.

    References1. Hodgkin AL, Huxley AF. A quantitative description of membrane current

    and its application to conduction and excitation in nerve. J Physiol 1952;

    117:500544.

    2. Bock G, Goode J. (Eds.) Novartis Foundation Symposium no 239,

    Complexity in Biological Information Processing, London: Wiley; 2001.

    3. Bock G, Goode J. (Eds.) Novartis Foundation Symposium no 247,

    In Silico simulation of biological processes. London: Wiley; 2002.

    4. Noble D, Rudy Y. Models of cardiac ventricular action potentials: iterative

    interaction between experiment and simulation. Phil Trans Roy Soc A

    2001;359:11271142.

    5. FitzHugh R. Thresholds and plateaus in the Hodgkin-Huxley nerve

    equations. J Gen Physiol 1960;43:867896.

    6. Weidmann S. Effect of current flow on the membrane potential of cardiac

    muscle. J Physiol 1951;115:227236.

    7. Weidmann S. Elektrophysiologie der Herzmuskelfaser. Bern: Huber;

    1956.

    8. Hutter OF, Noble D. Rectifying properties of heart muscle. Nature 1960;

    188:495.

    9. Carmeliet EE. Chloride ions and the membrane potential of Purkinje

    fibres. J Physiol 1961;156:375388.

    10. Hall AE, Hutter OF, Noble D. Current-voltage relations of Purkinje fibres

    in sodium-deficient solutions. J Physiol 1963;166:225240.

    11. Noble D. Cardiac action and pacemaker potentials based on the

    Hodgkin-Huxley equations. Nature 1960;188:495 497.

    12. Noble D. A modification of the Hodgkin-Huxley equations applicable to

    Purkinje fibre action and pacemaker potentials. J Physiol 1962;160:317

    352.

    13. Muzikant AL, Penland RC. Models for profiling the potential QT pro-

    longation risk of drugs. Current Opin Drug Disc Devel 2002;5:127135.

    14. Deck KA, Trautwein W. Ionic currents in cardiac excitation. Pflugers

    Archiv 1964;280:6580.

    15. Reuter H. The dependence of the slow inward current in Purkinje fibres

    on the extracellular calcium concentration. J Physiol 1967;192:479492.

    16. Noble D, Tsien RW. Outward membrane currents activated in the plateau

    range of potentials in cardiac Purkinje fibres. J Physiol 1969;200:205

    231.

    17. Sanguinetti MC, Jurkiewicz NK. Two components of cardiac delayed

    rectifier K current:differential sensitivity to block by class III antiar-

    rhythmic agents. J Gen Physiol 1990;96:195215.

    18. Noble D, Tsien RW. The kinetics and rectifier properties of the slow

    potassium current in cardiac Purkinje fibres. J Physiol 1968;195:185

    214.

    19. McAllister RE, Noble D, Tsien RW. Reconstruction of the electrical activity

    of cardiac Purkinje fibres. J Physiol 1975;251:159.

    20. DiFrancesco D. A new interpretation of the pace-maker current, iK2, in

    Purkinje fibres. J Physiol 1981;314:359376.

    21. DiFrancesco D, Noble D. Implications of the re-interpretation of iK2 for the

    modelling of the electrical activity of pacemaker tissues in the heart.

    In: Bouman LN, Jongsma H.J, eds. Cardiac Rate and Rhythm. The

    Hague, Boston, London: Martinus Nijhoff; 1982. pp. 93128.

    22. Noble D. The surprising heart: a review of recent progress in cardiac

    electrophysiology. J Physiol 1984;353:150.

    23. Beeler GW, Reuter H. Reconstruction of the action potential of ventricular

    myocardial fibres. J Physiol 1977;268:177210.

    24. DiFrancesco D, Noble D. A model of cardiac electrical activity incorpo-

    rating ionic pumps and concentration changes. Phil Trans Roy Soc B

    1985;307:353398.

    25. Chen FF-T, Vaughan-Jones RD, Clarke K, Noble D. Modelling my-

    ocardial ischaemia and reperfusion. Prog Biophys Mol Biol 1998;69:

    515538.

    26. Winslow RL, Greenstein JL, Tomaselli GF, ORourke B. Computational

    models of the failing myocyte: relating altered gene expression to cellular

    function. Phil Trans Roy Soc A 2001;359:11871200.

    27. Kimura J, Noma A, Irisawa H. Na-Ca exchange current in mammalian

    heart cells. Nature 1986;319:596597.

    28. Varghese A, Winslow RL. Dynamics of the calcium subsystem in cardiacPurkinje fibres. Physica D 1993;68:364386.

    29. Varghese A, Winslow RL. Dynamics of abnormal pacemaking activity in

    cardiac Purkinje fibres. J Theoret Biol 1994;168:407420.

    30. Winslow RL, Rice J, Jafri S, et al. Mechanisms of altered excitation-

    contraction coupling in canine tachycardia-induced heart failure, II Model

    studies. Circ Res 1999;84(5):571586.

    31. White JA, Guckenheimer J, Winslow RL. Spontaneous calcium release

    in ventricular myocytes: mechanisms and implications. Chaos 2002;

    (in press).

    Review articles

    1162 BioEssays 24.12

  • 8/8/2019 Noble2002-Modelling the Heart-Insights Failures and Progress

    9/9

    32. Noble D. Modelling the heart: from genes to cells to the whole organ.

    Science 2002;295:16781682.

    33. Hilgemann DW, Noble D. Excitation-contraction coupling and extracel-

    lular calcium transients in rabbit atrium: Reconstruction of basic cellular

    mechanisms. Proc Roy Soc B 1987;230:163205.

    34. Luo CH, Rudy Y. A model of the ventricular cardiac action potential:

    Depolarization, repolarization, and their interaction. Circ Res 1991;68:

    15011526.

    35. Noble D, Noble SJ, Bett GCL, Earm YE, Ho WK, So IS. The role of

    sodium-calcium exchange during the cardiac action potential. Ann NY

    Acad Sci 1991;639:334353.

    36. Luo CH, Rudy Y. A dynamic model of the cardiac ventricular action

    potential: I. Simulations of ionic currents and concentration changes.

    Circ Res 1994;74:10711096.

    37. Luo CH, Rudy Y. A dynamic model of the cardiac ventricular action

    potential: II. Afterdepolarizations, triggered activity and potentiation. Circ

    Res 1994;74:10971113.

    38. Noble D, Varghese A, Kohl P, Noble PJ. Improved guinea-pig ventricular

    cell model incorporating a diadic space, iKr & iKs, and length- & tension-

    dependent processes. Can J Cardiol 1998;14:123134.

    39. Hilgemann DW. Extracellular calcium transients and action potential

    configuration changes related to post-stimulatory potentiation in rabbit

    atrium. J Gen Physiol 1986;87:675706.

    40. Mitchell MR, Powell T, Terrar DA, Twist VA. The effects of ryanodine,

    EGTA and low-sodium on action potentials in rat and guinea-pig ventri-

    cular myocytes: evidence for two inward currents during the plateau.

    Brit J Pharmacol 1984;81:543550.

    41. Fabiato A. Calcium-induced release of calcium from the cardiac

    sarcoplasmic reticulum. Am J Physiol 1983;245:C1C14.

    42. Fabiato A. Time and calcium dependence of activation and inactivation

    of calcium-induced release of calcium from the sarcoplasmic reticulum

    of a skinned canine cardiac Purkinje cell. J Gen Physiol 1985;85:247

    298.

    43. Jafri MS, Rice JJ, Winslow RL. Cardiac Ca2 Dynamics: The Roles of

    Ryanodine Recptor Adaptation and Sarcoplasmic Reticulum Load.

    Biophys J 1998;74:11491168.

    44. Winslow RL, Scollan DF, Holmes A, Yung CK, Zhang J, Jafri MS.

    Electrophysiological modeling of cardiac ventricular function: from cell to

    organ. Ann Rev Biomed Eng 2000;2:119155.

    45. LeGuennec JY, Noble D. The effects of rapid perturbation of external

    sodium concentration at different moments of the action potential in

    guinea-pig ventricular myocytes. J Physiol 1994;478:493 504.

    46. Winslow RL, Varghese A, Noble D, Adlakha C, Hoythya A. Generation

    and propagation of triggered activity induced by spatially localised Na-K

    pump inhibition in atrial network models. Proc Roy Soc B 1993;254:

    5561.

    47. Earm YE, Noble D. A model of the single atrial cell: relation between

    calcium current and calcium release. Proc Roy Soc B 1990;240:

    8396.

    48. Shaw RM, Rudy Y. Electrophysiologic effects of acute myocardial

    ischemia: a mechanistic investigation of action potential conduction and

    conduction failure. Circ Res 1997;80:124138.

    49. Shaw RM, Rudy Y. Ionic mechanisms of propagation in cardiac tissue:

    roles of the sodium and L-type calcium currents during reduced ex-

    citability and decreased gap junction coupling. Circ Res 1997;81:727

    741.

    50. Shaw RM, Rudy Y. Electrophysiologic effects of acute myocardial

    ischemia: a theorectical study of altered cell excitability and action

    potential duration. Cardiovasc Res 1997;35:256272.

    51. Panfilov A, Holden A. Computer simulation of re-entry sources in

    myocardium in two and three dimensions. J Theoret Biol 1993;161:

    271285.

    52. Panfilov A, Keener J. Re-entry generation in anisotropic twisted

    myocardium. J Cardiovascular Electrophysiol 1993;4:412421.

    53. Nielsen PMF, LeGrice IJ, Smaill BH, Hunter PJ. A mathematical model of

    the geometry and fibrous structure of the heart. Am J Physiol 1991;29:

    H1365H1378.

    54. Winslow RL, Scollan D. 1997; Modelling normal and abnormal

    electrical activity in the canine ventricle: from single cells to whole heart.

    In XXXIII International Congress of the IUPS pp. L017.08. (abstract),

    St. Petersburg.

    55. Sands GB. 1998. Mathematical model of ventricular activation in an

    anatomically accurate deforming heart. In Department of Engineering

    Science, University of Auckland, Auckland

    56. Hunter PJ, Nash MP, Sands GB. Computational electromechanics of

    the heart. In: Panfilov A, Holden A, editors. Computational Biology of the

    Heart. Chichester: John Wiley & Sons; 1997. pp. 345407.

    57. Hunter PJ, Smaill BH. The analysis of cardiac function - a continuum

    approach. Prog Biophys Mol Biol 1998;52:101164.

    58. Smith NP, Kassab G. Analysis of coronary blood flow interaction with

    myocardial mechanics based on anatomical models. Phil Trans Roy Soc

    Lond A 2001;359:12511262.

    59. LeGrice I, Hunter P, Young A, Smaill B. The architecture of the heart: a

    data-based model. Phil Trans Roy Soc Lond A 2001;359:12171232.

    60. Skouibine K, Trayanova N, Moore PA. Numerically efficient method for

    simulation of defibrillation in an active bidomain sheet of myocardium.

    Mathematical Biosciences 2000;166:85 100.

    61. Trayanova N, Eason J. Shock-induced arrhythmogenesis in the

    myocardium. Chaos 2002;12:962972.

    62. Trayanova N, Eason J, Aguel F. Computer simulations of cardiac

    defibrillation: A look inside the heart. Computing and Visualization in

    Science 2002; (in press).

    63. Boyd CAR, Noble D. The Logic of Life. Oxford: OUP. 1993.

    64. Noble D. Biological Computation. In Encyclopedia of Life Sciences,

    http://www.els.net, London: Nature Publishing Group; 2002.

    65. Noble D. The rise of computational biology. Nature Rev Mol Cell Biol

    2002;3:460463.

    66. Hunter PJ, Kohl P, Noble D. Integrative models of the heart: achieve-

    ments and limitations. Phil Trans Roy Soc Lond A 2001;359:10491054.

    67. Clancy CE, Rudy Y. Linking a genetic defect to its cellular phenotype in a

    cardiac arrhythmia. Nature 1999;400:566569.

    68. Noble D, Noble PJ. Reconstruction of cellular mechanisms of genetically-

    based arrhythmias. J Physiol 1999;518:2P3P.

    69. Noble PJ, Noble D. Reconstruction of the cellular mechanisms of cardiac

    arrhythmias triggered by early after-depolarizations. Japanese Journal of

    Electrocardiology 2000;20(suppl 3):1519.

    70. Wehrens XHT, Abriel H, Cabo C, Benhorin MD, Kass RS. Arrhythmognic

    mechanism of an LQT-3 mutation of the human heart Na channel

    a-subunit. Circulation 2000;102:584 590.

    71. Nuyens D, Stengl M, et al. Abrupt rate accelerations or premature beats

    cause life-threatening arrhythmias in mice with long QT syndrome.

    Nature Medicine 2001;7:10211027.

    Review articles

    BioEssays 24.12 1163