Improving the Stability of Cardiac Mechanical Simulations

9
0018-9294 (c) 2013 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission. See http://www.ieee.org/publications_standards/publications/rights/index.html for more information. This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI 10.1109/TBME.2014.2373399, IEEE Transactions on Biomedical Engineering IEEE TRANSACTIONS ON BIOMEDICAL ENGINEERING VOL. , NO. , MONTH 0000 1 Improving the stability of cardiac mechanical simulations Sander Land, Steven A. Niederer, Pablo Lamata, Nicolas P. Smith Abstract—In the field of cardiac modelling, the mechanical action of the heart is often simulated using finite element methods. These simulations are becoming increasingly challenging as the computational domain is customized to a patient’s anatomy, within which large heterogeneous tension gradients are generated via biophysical cell models which drive simulations of the cardiac pump cycle. The convergence of nonlinear solvers in simulations of large deformation mechanics depends on many factors. When extreme stress or irregular deformations are modelled, commonly used numerical methods can often fail to find a solution, which can prevent investigation of interesting parameter variations or use of models in a clinical context with high standards for robustness. This article outlines a novel numerical method that is straightforward to implement and which significantly improves the stability of these simulations. The method involves adding a compressibility penalty to the standard incompressible formula- tion of large deformation mechanics. We compare the method’s performance when used with both a direct discretization of the equations for incompressible solid mechanics, as well as the formulation based on an isochoric/deviatoric split of the deformation gradient. The addition of this penalty decreases the tendency for solutions to deviate from the incompressibility constraint, and significantly improves the ability of the Newton solver to find a solution. Additionally our method maintains the expected order of convergence under mesh refinement, has nearly identical solutions for the pressure-volume relations, and stabilizes the solver to allow challenging simulations of both diastolic and systolic function on personalized patient geometries. Index Terms—solid mechanics, incompressibility, nonlinear solvers, cardiac mechanics I. I NTRODUCTION Finite element methods are increasingly used to study the mechanics of the heart. Studies in this area range from detailed electromechanical simulations with the goal of improving our understanding of physiology [1], [2], [3], genetically modified animals [4], [5], disease [6], [7], [8], and clinical intervention [9]. As the range of applications for these models is expanding, robustness of simulations over a large number of different cardiac geometries and physiological perturbations is becoming increasingly important. Manuscript received ...; revised ... This work was supported by the BBSRC (BB/J017272/1) and the EPSRC (EP/F043929/1 and EP/G007527/2). P. Lamata holds a Sir Henry Dale Fellowship jointly funded by the Wellcome Trust and the Royal Society (099973/Z/12/Z). The authors acknowledge financial support from the De- partment of Health via the National Institute for Health Research (NIHR) comprehensive Biomedical Research Centre award to Guy’s & St Thomas’ NHS Foundation Trust in partnership with King’s College London and King’s College Hospital NHS Foundation Trust S. Land, S.A. Niederer, P. Lamata and N.P. Smith are with the Biomedical Engineering Department, King’s College London, London SE17EH, UK. P. Lamata is with the Department of Computer Science, University of Oxford, Oxford OX13QD, UK. N.P. Smith is with the Department of Engineer- ing, The University of Auckland, Auckland, New Zealand. Email address: [email protected] Tel.: +649 923 7924 Fax: +649 373 7464 A problem that frequently arises in solving these large deformation mechanics problems is that the nonlinear solution processes do not guarantee convergence to a solution of the set of governing equations [10]. As the number of degrees of freedom in these problems is typically in the tens to hundreds of thousands, it is difficult to uniquely identify the cause of such solver nonconvergence in a given context. These con- vergence problems can prevent researchers from investigating the effect of a large range of perturbations to the heart, as the more extreme (and often more interesting) variations are less likely to be solved. Furthermore, the convergence behavior of a solver for a biophysically detailed simulation of cardiac mechanics is sensitive to many other factors, including the computational mesh, choice of spatial or temporal numerical discretization, model parameters, boundary conditions, and many aspects of the numerical solver such as the matrix solver and line search algorithm. Solver convergence issues have become even more relevant in the context of personalizing cardiac geometries for mechan- ical simulations [6], [11], where image processing and mesh fitting procedures now significantly influence the ability of a nonlinear solver to find a solution to the computational model. In particular, matching the patient’s anatomy more accurately has been shown to result in a more challenging computational problem, and achieving a good balance between accuracy and solver stability is a difficult problem. Furthermore, although mesh quality metrics can be maximized during mesh gener- ation, they can not be robustly used to determine the solver performance [10]. Robust solver convergence when using these models is also important in the context of the parameter estimation required to further personalize these cardiac models [12], as the methods used to run the estimation procedures assume the simulations can be evaluated for a wide range of parameters. There are a large number of numerical schemes that have been proposed for the numerical solution of the solid me- chanics equations in the past fifty years. Many of these schemes are designed to overcome the difficulties of modelling incompressible materials. Two common approaches are the use of penalty methods and Lagrange multiplier methods. Penalty methods approximate fully incompressible materials as nearly incompressible, penalizing changes in volume causing with a large increase in strain energy. The compressibility of the material decreases as the penalty term is increased, but the resulting system of equations becomes numerically more dif- ficult to solve. Several numerical schemes have been devised to alleviate these problems, including updating schemes based on augmented Lagrangians. For more information on these schemes and other method based on specific finite element im-

Transcript of Improving the Stability of Cardiac Mechanical Simulations

Page 1: Improving the Stability of Cardiac Mechanical Simulations

0018-9294 (c) 2013 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission. Seehttp://www.ieee.org/publications_standards/publications/rights/index.html for more information.

This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI10.1109/TBME.2014.2373399, IEEE Transactions on Biomedical Engineering

IEEE TRANSACTIONS ON BIOMEDICAL ENGINEERING VOL. , NO. , MONTH 0000 1

Improving the stability of cardiac mechanicalsimulations

Sander Land, Steven A. Niederer, Pablo Lamata, Nicolas P. Smith

Abstract—In the field of cardiac modelling, the mechanicalaction of the heart is often simulated using finite element methods.These simulations are becoming increasingly challenging as thecomputational domain is customized to a patient’s anatomy,within which large heterogeneous tension gradients are generatedvia biophysical cell models which drive simulations of the cardiacpump cycle. The convergence of nonlinear solvers in simulationsof large deformation mechanics depends on many factors. Whenextreme stress or irregular deformations are modelled, commonlyused numerical methods can often fail to find a solution, whichcan prevent investigation of interesting parameter variations oruse of models in a clinical context with high standards forrobustness. This article outlines a novel numerical method that isstraightforward to implement and which significantly improvesthe stability of these simulations. The method involves adding acompressibility penalty to the standard incompressible formula-tion of large deformation mechanics. We compare the method’sperformance when used with both a direct discretization ofthe equations for incompressible solid mechanics, as well asthe formulation based on an isochoric/deviatoric split of thedeformation gradient. The addition of this penalty decreasesthe tendency for solutions to deviate from the incompressibilityconstraint, and significantly improves the ability of the Newtonsolver to find a solution. Additionally our method maintainsthe expected order of convergence under mesh refinement, hasnearly identical solutions for the pressure-volume relations, andstabilizes the solver to allow challenging simulations of bothdiastolic and systolic function on personalized patient geometries.Index Terms—solid mechanics, incompressibility, nonlinearsolvers, cardiac mechanics

I. INTRODUCTION

Finite element methods are increasingly used to study themechanics of the heart. Studies in this area range from detailedelectromechanical simulations with the goal of improvingour understanding of physiology [1], [2], [3], geneticallymodified animals [4], [5], disease [6], [7], [8], and clinicalintervention [9]. As the range of applications for these modelsis expanding, robustness of simulations over a large number ofdifferent cardiac geometries and physiological perturbations isbecoming increasingly important.

Manuscript received ...; revised ...This work was supported by the BBSRC (BB/J017272/1) and the EPSRC

(EP/F043929/1 and EP/G007527/2). P. Lamata holds a Sir Henry DaleFellowship jointly funded by the Wellcome Trust and the Royal Society(099973/Z/12/Z). The authors acknowledge financial support from the De-partment of Health via the National Institute for Health Research (NIHR)comprehensive Biomedical Research Centre award to Guy’s & St Thomas’NHS Foundation Trust in partnership with King’s College London and King’sCollege Hospital NHS Foundation Trust

S. Land, S.A. Niederer, P. Lamata and N.P. Smith are with the BiomedicalEngineering Department, King’s College London, London SE17EH, UK. P.Lamata is with the Department of Computer Science, University of Oxford,Oxford OX13QD, UK. N.P. Smith is with the Department of Engineer-ing, The University of Auckland, Auckland, New Zealand. Email address:[email protected] Tel.: +649 923 7924 Fax: +649 373 7464

A problem that frequently arises in solving these largedeformation mechanics problems is that the nonlinear solutionprocesses do not guarantee convergence to a solution of theset of governing equations [10]. As the number of degrees offreedom in these problems is typically in the tens to hundredsof thousands, it is difficult to uniquely identify the cause ofsuch solver nonconvergence in a given context. These con-vergence problems can prevent researchers from investigatingthe effect of a large range of perturbations to the heart, as themore extreme (and often more interesting) variations are lesslikely to be solved. Furthermore, the convergence behaviorof a solver for a biophysically detailed simulation of cardiacmechanics is sensitive to many other factors, including thecomputational mesh, choice of spatial or temporal numericaldiscretization, model parameters, boundary conditions, andmany aspects of the numerical solver such as the matrix solverand line search algorithm.

Solver convergence issues have become even more relevantin the context of personalizing cardiac geometries for mechan-ical simulations [6], [11], where image processing and meshfitting procedures now significantly influence the ability of anonlinear solver to find a solution to the computational model.In particular, matching the patient’s anatomy more accuratelyhas been shown to result in a more challenging computationalproblem, and achieving a good balance between accuracy andsolver stability is a difficult problem. Furthermore, althoughmesh quality metrics can be maximized during mesh gener-ation, they can not be robustly used to determine the solverperformance [10]. Robust solver convergence when using thesemodels is also important in the context of the parameterestimation required to further personalize these cardiac models[12], as the methods used to run the estimation proceduresassume the simulations can be evaluated for a wide range ofparameters.

There are a large number of numerical schemes that havebeen proposed for the numerical solution of the solid me-chanics equations in the past fifty years. Many of theseschemes are designed to overcome the difficulties of modellingincompressible materials. Two common approaches are the useof penalty methods and Lagrange multiplier methods. Penaltymethods approximate fully incompressible materials as nearlyincompressible, penalizing changes in volume causing witha large increase in strain energy. The compressibility of thematerial decreases as the penalty term is increased, but theresulting system of equations becomes numerically more dif-ficult to solve. Several numerical schemes have been devisedto alleviate these problems, including updating schemes basedon augmented Lagrangians. For more information on theseschemes and other method based on specific finite element im-

Page 2: Improving the Stability of Cardiac Mechanical Simulations

0018-9294 (c) 2013 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission. Seehttp://www.ieee.org/publications_standards/publications/rights/index.html for more information.

This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI10.1109/TBME.2014.2373399, IEEE Transactions on Biomedical Engineering

IEEE TRANSACTIONS ON BIOMEDICAL ENGINEERING VOL. , NO. , MONTH 0000 2

plementations see e.g. [13], [14], [15] and references therein.A technique which can be applied in both nearly- and fully in-compressible modelling, are schemes based on a multiplicativesplit of the deformation gradient into isochoric and deviatoriccomponents, which can be traced back to early work of Flory[16]. Within these schemes, and indeed for the material itself,the stress response is assumed to depend only on the isochoric(i.e. volume-preserving) part of the deformation. Although inthe formulation of fully incompressible solid mechanics, thereis no volume change in the exact solution, this assumptionaffects the numerical solution process, making it generallymore likely to converge to a solution [17]. These methodshave also been used in cardiac modelling, primarily in thecontext of simulating diastolic inflation of the ventricles [18],[19], For a more detailed review of these methods and theirtheoretical foundation based on the Hu-Washizu principle,see e.g. [20], [21], [17]. Although this isochoric/deviatoricformulation is known to be particularly robust [17], a directdiscretization of the equations for solid mechanics, withoutusing an alternative stabilizing formulation, remains the mostcommon approach in cardiac modelling [22], [23], [24], [25],[26], [27], [28], [29]. The fact that with very few exceptions[30], this includes nearly all detailed whole-cycle electrome-chanical models suggests that the implementation challenge ofusing these schemes, combined with the uncertainties of theireffectiveness in solving whole-cycle cardiac simulations, mayrepresent a significant hurdle to their adoption.

Given the lack of application of these more advancedschemes to complex cardiac simulation, and the apparentimplementation barriers, the objectives of this paper aretwofold. Firstly, we investigate the performance of the directand isochoric/deviatoric formulations of the solid mechanicsequations for a range of problems in cardiac mechanics,including both extreme inflation and simulations of the cardiaccycle. Secondly, we present an alternative scheme which isboth simple and practical, and which significantly improves thestability and accuracy of cardiac simulations. This adjustmentis straightforward to implement in a code using the direct finiteelement discretization of the solid mechanics equations, andcalculation of the term has a negligible additional compu-tational cost. The scheme can be applied to both the directand isochoric/deviatoric formulations, and we compare theeffect of this scheme on both methods. Below we presentthe mathematical background of our proposed novel numericalscheme. We then perform a detailed analysis on a simple testproblem, to show the convergence under mesh refinement ofthe different numerical schemes. Finally, we analyze the stabil-ity and computational performance of the different methods ona range of practical problems of personalized cardiac modelsof human patients.

II. METHODS AND RESULTS

Cardiac tissue is typically modelled as an incompressiblematerial [22], [24], [9], [26]. Although cardiac cells do notchange volume over the course of a beat, estimates of tissuevolume change during the cardiac cycle range from 2− 15%[31]. As these volume changes are driven by changes in

perfusion, we approximate cardiac tissue as a fully incom-pressible material, in the absence of a perfusion componentin our model. In large deformation mechanics, the governingequations are given by:

∀i :∑N

ddXN

(∑M

FiMTMN

)= 0 (1)

Where F is the deformation gradient (Fij = dxi

dXj), and

coordinates in the undeformed and deformed configuration aredenoted by X and x respectively. In equation 1, T is thesecond Piola-Kirchhoff stress, given by the derivative of thestrain energy function W :

T = 2dWdC

=dWdE

(2)

Where E = 12 (C − I),C = FTF are the Lagrangian and

right Cauchy-Green strain tensors respectively. For the purposeof demonstration we will use the exponential strain-energyfunction proposed by Guccione et al. [32] throughout thismanuscript:

Wg(E) =C1

2(eQ − 1) where (3)

Q = C2E2ff + C3(E2

ss + E2nn + 2E2

sn) + 2C4(E2fs + E2

fn)

Where f, s, n have the conventional definition of ‘fiber direc-tion’, ‘sheet direction’ and ‘normal direction’ in local tissuemicrostructure coordinates for cardiac simulations [33].There are two main approaches to incompressibility in cardiacmechanics: the Lagrange multiplier method, and the penaltymethod. These are given by the strain energy functions:

Wlagrange = Wg(E)− p(J − 1) with constraint J = 1 (4)

Wpenalty = Wg(E) +κ

2(J − 1)2 (5)

Where p is the hydrostatic pressure and J = det F. Inaddition, we investigate an alternative discretization, as usedby Goktepe et al. and Wang et al. among others [34], [18],[19]. This scheme defines the isochoric component of thedeformation gradient as F = J−1/3F, along with isochoricstrain tensors C,E, and uses these to define a strain energyfunction independent of changes in volume:

Wiso = Wg(E)− p(J − 1) with constraint J = 1 (6)

Analogous variations of penalty methods (equation 5) alsoexists, but are susceptible to non-physical behaviour [35] andwill not be considered here. We focus on the often adoptedchoice in cardiac mechanics where equation 1 is solved usingtricubic Lagrange elements and the constraint J = 1 issolved using trilinear elements where applicable [22], [4].Specifically, the weak form of the incompressibility constraintJ = 1 in terms of basis functions φj is:

∀j∫X

(J − 1)φj dV = 0 (7)

In analyzing situations where the Newton solver does notconverge to a solution, we observed that the local change involume J can be significantly different from unity while stillobeying equation 7, and observed non-physical trial solutions

Page 3: Improving the Stability of Cardiac Mechanical Simulations

0018-9294 (c) 2013 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission. Seehttp://www.ieee.org/publications_standards/publications/rights/index.html for more information.

This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI10.1109/TBME.2014.2373399, IEEE Transactions on Biomedical Engineering

IEEE TRANSACTIONS ON BIOMEDICAL ENGINEERING VOL. , NO. , MONTH 0000 3

where J < 0. Based on this observation, we introduce twonovel schemes, based on adding a compressibility penalty (asin equation 5) to the Lagrange multiplier schemes.

W stablagrange = Wg(E)− p(J − 1) +

κ

2(J − 1)2 (8)

W stabiso = Wg(E)− p(J − 1) +

κ

2(J − 1)2 (9)

Where the incompressibility constraint J = 1 applies in bothschemes. The deformation (and thus J) is represented byhigher order finite elements, while p is represented by trilinearelements. Thus, the addition of a higher order incompress-ibility penalty term means that the deformation is expectedto more accurately obey the incompressibility constraint. Inaddition, the strain energy in equation 8 is similar to thatused in augmented Lagrange schemes [15], which iterativelyupdate p at each Gauss point and use sub-iterations to achieveincompressibility. However, these schemes do not solve the in-compressibility constraint J = 1 directly, but instead representa variation of the strain energy in equation 5. The electronicsupplement includes derivations of the Piola-Kirchhoff stresstensors for the schemes we compare, as well as mathematicaldetails.

As discussed in the introduction, there are many othernumerical schemes for solid mechanics. In this paper, we limitour investigation to methods that are commonly used in cardiacmechanics and our proposed novel variations on them, i.e.the five strain energy functions given in this section. We alsolimit the investigation of the penalty method to κ = 1000 kPa,which limits the difference in volume to approximately 10%compared to fully incompressible schemes in the physiolog-ical range of pressure and stiffness. The following sectionsshow the effect of these different numerical methods on theconvergence under mesh refinement and solver stability ofmechanical simulations.

A. Tests on a cylinder problemIn this section we present an analysis of the conver-

gence behaviour under mesh refinement of the five differ-ent schemes described in the previous section (the directand isochoric/deviatoric schemes, both with and without theadditional stabilizing term, and a penalty method). For thispurpose, we consider a simple test problem by inflating athin cylinder of radius 30 mm and thickness 3 mm to typicalend diastolic pressure of Pcav = 1 kPa. This test problemallows us to unambiguously analyze differences between thevarious computational schemes, without the complicating fac-tors inherited from methods used in fitting the computationaldomain to physiological data, while still retaining some ofthe properties of the deformations seen in simulations ofcardiac mechanics. The symmetry of the problem also makesit more tractable, as mesh refinement in two dimensions issufficient to study convergence. For simplicity we define thematerial to be isotropic by using C2 = C3 = C4 to eliminatethe dependence on the fiber direction, specifically we useC1 = 1 kPa, C2 = C3 = C4 = 5 to approximately doublevolume at the inflation pressure. We use a small load incrementto maximize the change of convergence for the first solversteps, incrementing pressure by 0.01 kPa each step.

For this problem we consider the following three errormetrics. Firstly, we consider the mean error in the deformationin Cartesian coordinates:

ErrX = meanx∈A

‖xconv − x‖ (10)

Where x is the node position, and the set of points A to becompared are all the nodes in the least refined mesh, includingnodes internal to elements. We use a further refined solutionwith 128 × 8 elements to determine the converged solutionxconv. Secondly, we consider the error as measured by thetotal strain energy (in Joules):

ErrE =

∫W dV −

∫Wconv dV (11)

Thirdly, we consider the error in solving the incompressibilityconstraint, per unit volume, which measures how well thesolution found obeys the incompressibility constraint:

ErrI =

√∫mesh

(det F− 1)2 dV/∫

meshdV (12)

For the stabilization terms we initially set the compress-ibility penalty parameter to κ = 5Pcav. The dependence ofκ on cavity pressure compensates well for the tendency forthe solution to show greater deviation from the theoreticallyincompressible cases in larger deformations due to greatercavity pressures being applied. We will consider the choiceof κ in more detail later in this paper. Figure 2 showsconvergence under mesh refinement for all four numericalschemes on the meshes shown in Figure 1. These plots showthat the proposed stabilization preserves the solution in termsof pressure-volume relationship and has identical convergencebehaviour compared to standard methods, while giving amore accurate solution for the incompressibility constraint.Table I shows differences between the methods are on theorder of 10−6 mm, with similar differences resulting fromusing the isochoric/deviatoric discretization over the direct dis-cretization, as resulting from the introduction of a stabilizingterm. Additional benchmark problems confirming the lack oflocking across different schemes are shown in the electronicsupplement.

B. Investigating nonconvergence of the Newton solver

As outlined in the methods section, the addition of acompressibility penalty to the stress tensor was motivated bythe observation of a negative det F in the Newton iterationprocess. However, we have found that mesh refinement canresult in a the incompressibility equation being solved moreaccurately, yet still resulting in a negative det F in the Newtonsearch direction and non-convergence of the solver, at nearlyidentical pressures. In this section we investigate a number ofalternative explanations for solver non-convergence, includingnumerical roundoff, the choice of initial solution, and the prop-erties of the Jacobian at the point of solver non-convergence.We use the 4× 1 mesh for efficiency reasons, as all problemsshow similar behaviour. We define the vector of unknowns asx in solving the system of equations ‖r(x)‖ < Nrreltol whereN is the number of nodes and our default rreltol = 10−9. For

Page 4: Improving the Stability of Cardiac Mechanical Simulations

0018-9294 (c) 2013 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission. Seehttp://www.ieee.org/publications_standards/publications/rights/index.html for more information.

This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI10.1109/TBME.2014.2373399, IEEE Transactions on Biomedical Engineering

IEEE TRANSACTIONS ON BIOMEDICAL ENGINEERING VOL. , NO. , MONTH 0000 4

Fig. 1. Meshes used for the cylinder problem. Since the problem is perfectly symmetric, we only refine the mesh circumferentially and radially as shownin panel (a), indicating the level of refinement as ‘number of elements circumferentially × number of elements radially’.

Fig. 2. Convergence results for the cylinder problem. Our refinement study shows fourth-order convergence for deformation (Panel A, Equation 10), andshows second-order convergence for the incompressibility constraint (Panel B, Equation 12). Panel B also shows that our proposed method can significantlyreduce error in solving this equation. Results for all methods are compared to a refined solution computed with the same method, while table I showsdifferences between methods for these refined solutions. Panels C shows error in total strain energy convergence, while Panel D confirms convergence toidentical total strain energy for all fully incompressible schemes.

Method D DS I IS PDirect discretization (D) - 6.8 · 10−7 1.9 · 10−6 1.8 · 10−6 5.4 · 10−3

... with stabilization (DS) - 1.4 · 10−6 1.3 · 10−6 5.4 · 10−3

Isochoric/deviatoric discretization (I) - 1.4 · 10−7 5.4 · 10−3

... with stabilization (IS) - 5.4 · 10−3

Penalty (P) -

TABLE IDIFFERENCES IN DEFORMATION FOR THE DIFFERENT NUMERICAL METHOD. RESULTS SHOW MEAN DIFFERENCES IN DEFORMATION BETWEEN THE

DIFFERENT METHODS, AS MEASURED BY THE MEAN DIFFERENCE BETWEEN NODES IN THE MOST HIGHLY REFINED MESH (128 CIRCUMFERENTIALELEMENTS), MEASURED IN MM. TABLE ENTRIES AT THE TOP ARE ABBREVIATIONS FOR THE DIFFERENT NUMERICAL METHODS AS INDICATED IN

PARENTHESIS AT THE START OF EACH ROW. AS THE PENALTY METHOD HAS A DIFFERENT EXACT SOLUTION, DIFFERENCES WITH THIS SCHEME ARESIGNIFICANTLY LARGER, LEADING TO IDENTICAL ENTRIES IN THE LAST COLUMN AT THIS LEVEL OF PRECISION.

this choice of tolerance, inflation pressure is 48 kPa for thedirect discretization, increasing to > 100 kPa when adding astabilization term.

Firstly, we test whether loss of numerical precision mayprevent the residual r from decreasing sufficiently. Increasingrreltol also results in non-convergence of the Newton solver,with an identical inflation pressure for rreltol = 10−6 and sur-prisingly a lower inflation pressure of 35 kPa for rreltol = 10−3.However, decreasing rreltol to e.g. 10−12 results in earlierfailure (< 2 kPa) in both the non-stabilized and stabilizedmethod. This shows that the solver convergence problems arenot caused purely by numerical roundoff for our typical choiceof rreltol = 10−9, but reflect a more fundamental problem.More detailed results for the dependence of solver convergence

on this tolerance as well as mesh refinement and discretizationmethod are shown in the electronic supplement.

Secondly, we tested the hypothesis that the choice of ini-tial solution is the typical cause of solver non-convergence.Running with the direct discretization to the point of non-convergence, and using the last iteration as a starting pointfor the stabilized method (κ = 1 kPa) results in convergencefor that load increment. Trying to re-solve the same loadincrement without stabilization, using the solution found withthe stabilized method, once again results in non-convergence,even though the initial guess for the solution is close to theone being sought. Similarly, switching back at a later loadincrement results in non-convergence. Thus, it appears unlikelythat finding a better initial solution is a practical solution for

Page 5: Improving the Stability of Cardiac Mechanical Simulations

0018-9294 (c) 2013 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission. Seehttp://www.ieee.org/publications_standards/publications/rights/index.html for more information.

This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI10.1109/TBME.2014.2373399, IEEE Transactions on Biomedical Engineering

IEEE TRANSACTIONS ON BIOMEDICAL ENGINEERING VOL. , NO. , MONTH 0000 5

solver nonconvergence.Finally, we have tested the properties of the nonlinear

system throughout the simulation by analyzing the eigenvaluesof the Jacobian matrix. The Jacobian matrix is typicallypositive definite, i.e. all eigenvalues are positive, and a lossof this property is associated with instability, as the Jacobianmatrix becomes (nearly) singular every time an eigenvaluecrosses from positive to negative. This presence of a near-zero eigenvalue in the Jacobian matrix A leads a large solverstep ∆x = A−1r, with trial solutions that are generally non-physical. Small eigenvalues can also also lead to a large con-dition number (largest eigenvalue/smallest eigenvalue) whichis associated with loss of numerical precision. To test theconnection between the eigenvalues and solver stability inpractice, we have explicitly determined the eigenvalues of theJacobian matrix each time the Jacobian matrix is calculated, toinvestigate which discretization schemes show negative eigen-values. Figure II-B shows the minimal eigenvalues throughoutthe solution process for the different discretization schemes.These results show that for the direct discretization, zerocrossings appear after 26.5 kPa inflation, with a risk for eigen-values close to zero and subsequent solver nonconvergence.These zero crossings also explain the irregular dependence onload increment and convergence tolerance resulting in inflationpressures between 35-48 kPa, and the fact that use of a smallerload increment can result in a lower inflation pressure. For thestabilized method, as well as the isochoric/deviatoric approachand the penalty method, no zero crossings appear, and solverperformance is more consistent. However, for higher pressuresthe Newton solver eventually fails, possibly due to the highcondition number (≈ 109), which is associated with lossof numerical precision. The isochoric/deviatoric discretizationshows especially high condition numbers, resulting in earlierfailure for this particular problem. The penalty method showsincreased stability at higher pressures, with the first load incre-ments being most difficult to solve in practice, but eventuallyfails at 790 kPa inflation.

Testing the dependence of solver convergence on the loadincrement (from 0.01 to 1 kPa) and choice of Newton solver(modified or not) confirms these observations. The directdiscretization shows less predictable results, as explained bythe chance of solver failure at each zero crossing of aneigenvalue, whereas the other methods are far more consistent.Furthermore, increasing the load increment makes the penaltymethod fail more often. Detailed results for these simulationsare presented in the electronic supplement.

In conclusion, solver non-convergence when using the directdiscretization appears to be related to a nearly singular Jaco-bian, which is associated with local extrema, saddle points, andbifurcations. The extra compressibility penalty term tends toprevent the Jacobian from becoming singular, thus improvingconvergence of the nonlinear solver. The improvement weare reporting may be related to improving the ellipticity andconvexity of the strain energy function, conditions which areimportant for preventing unnatural deformations. Specifically,the deformation block of the Jacobian being positive definite isimplied by the ‘strong ellipticity’ condition. Although the Guc-cione law has been shown to be locally convex for any positive

parameters [36], it is not strongly elliptic for anisotropic three-dimensional problems. Specifically, this important propertycan be violated under certain types of deformations involvingfiber compression in exponential strain energy laws [37] andwhen active contraction is added [25].

C. Personalized biventricular simulations

Our second test case relates to one of the primary motivationfor the stability improvement: allowing researchers to more ro-bustly simulate mechanics on personalized patient geometries.For this purpose, we have created twenty personalized biven-tricular cardiac meshes (shown in Figure 4) which representan accurate fit to a varied set of patient anatomies. The mesheswere generated using our mesh personalization tool [38] whichwas recently made available as a web service [39].

As a test for numerical stability, each of these twentyventricles is inflated to a left-ventricular pressure of Pcav =100 kPa, while maintaining right-ventricular pressure at 1

3 ofthe left ventricular pressure. For comparison, typical left-ventricular end-diastolic pressures are around 1 kPa [40]. Asin the previous section, we initially test three different settingsfor κ for the stabilized method, as in the cylinder problem. Weuse the constitutive parameters C1 = 4 kPa, C2 = 10, C3 =5, C4 = 2.5 based on typical volume changes at end-diastolicpressures. We set the fiber direction ranging from −60 degreesat the epicardium to +80 degrees at the endocardium.

Table II shows the results for these test runs, indicatingthe number of simulations which inflate to typical end-diastolic pressure, and those that succeed to extreme inflationof 100 kPa. As in the cylinder test, we use κ = 5Pcav.Using the direct method, none of the meshes reach extremeinflation. After adding the stabilization term with κ = 5Pcav,18/20 cases reach 100 kPa, while the other two meshes stillfail to reach 1 kPa inflation. Using the isochoric/deviatoricdiscretization, solver convergence is much improved, and onlyone case remains problematic. This is the same case thatshows worst-case behaviour in all methods, which suggeststhat generation of the computational mesh is still an importantfactor regardless of the numerical scheme used. Nevertheless,the addition of the stabilization term in the isochoric/deviatoricdiscretization roughly triples the inflation pressure for thiscase.

Comparing the solutions for the incompressible schemesreveals that differences are small. The left-ventricular cavityvolume when adding the stabilization term is approximately0.02% lower compared to the direct method without stabiliza-tion. Differences with the isochoric/deviatoric discretizationare larger, with volume almost 0.2% higher without a stabi-lization term, and 0.13% higher with this term included. Thepenalty method shows larger differences in volume which arevery load dependent. Although small at < 5 kPa, solutionsdiverge to a volume of approximately 50% higher than fullyincompressible schemes at 100 kPa.

Finally, these twenty cases also give us the opportunity tolook at the effects of the free parameter κ in more detail.So far we have only used the setting κ = 5Pcav, whichwas found to work well in practice. To explore the choice

Page 6: Improving the Stability of Cardiac Mechanical Simulations

0018-9294 (c) 2013 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission. Seehttp://www.ieee.org/publications_standards/publications/rights/index.html for more information.

This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI10.1109/TBME.2014.2373399, IEEE Transactions on Biomedical Engineering

IEEE TRANSACTIONS ON BIOMEDICAL ENGINEERING VOL. , NO. , MONTH 0000 6

Fig. 3. Eigenvalue analysis on the cylinder problem. Panel A shows the minimum absolute eigenvalue decreasing for higher pressures, leading to veryhigh condition numbers (Panel B). For the direct discretization without stabilization, the erratic values are caused by eigenvalues crossing zero (Panel C),which does not happen in the stabilized methods or isochoric/deviatoric approach.

Fig. 4. Biventricular meshes used for our test case. Twenty 144 element personalized geometries generated from MRI data of patients.

of stabilization parameter in more detail and determine itseffects on stability and computational cost, we run all 20biventricular cases with κ = κc + κp Pcav. Variations in boththe constant factor κc and the pressure-dependent factor κpare tested, using the values 0, 1, 5, 25, 125 for a total of 1000simulations. Results for inflation to 1 kPa are shown in Figure5. Most cases show a similar pattern, where the computationalcost increases significantly for higher κc, and less so forincreasing values of κp. A few especially challenging casesfail to solve except when using a large stabilization term orusing the deviatoric/isochoric formulation.

D. Whole-cycle biventricular simulations

In this section we present simulations of the full cardiaccycle on the same twenty biventricular patient geometries.Simulating a contracting heart adds a wider range of de-formations to our tests, and thus will better distinguish thepractical performance of the different numerical methods. Thewhole-cycle model with phase transitions is solved monolith-ically with the solid mechanics equations, as in our previ-ous work [3]. Three-element windkessel parameters for theejection phase were prescribed as R = 1.06 mmHg s/ml,C = 1.9 mmHg s/ml, Z = 0.062 ml/mmHg based on West-erhof et al. [41], with right-ventricular parameters set toR′ = R/3, Z ′ = Z/3, C ′ = 3C based on typical pressuredifferences between the ventricles. For active tension Ta weuse the model from Niederer et al. [8] based on work by

Kerckhoffs et al. [42], given by

Fiso = max (0, tanh(a6 − (λ− a7))) (13)

Ftwitch =

{tanh2(t/tr) tanh2((te − t)/td) t < te0 t ≥ te

(14)

te = b(λ− ld) (15)tr = tr0 + a4(1− Fiso) (16)Ta = T0FisoFtwitch (17)

T′ = T + TaffT (18)

Where tr0 = 80 ms, a4 = 650 ms, td = 90 ms, b = 1 ms, ld =−500, a6 = 6, a7 = 0.7, T0 = 180 kPa, and f is a unit vectorin the fibre direction. The stress tensor T′ including activetension is used instead of T in Equation 1. As is usual, λrepresents the local change in length in the fibre direction,given by λ = ‖Ff‖ =

√2Eff + 1.

We tested two different electrical activation patterns: firstly,a more physiological homogeneous activation, and secondly,a heterogeneous activation pattern which starts at the basalpart of the septum and spreads homogeneously with a con-duction velocity of 0.75 mm/ms, representative of pathologicalcases such as left bundle branch block. Results are shownin Table III. Notably, unlike in passive inflation tests, theisochoric/deviatoric formulation performs much worse than thedirect discretization in these cases without additional stabiliza-tion. The ejection and isovolumetric relaxation phase appearsespecially difficult to solve using this numerical method. Forboth formulations, the pressure-dependent penalty, although

Page 7: Improving the Stability of Cardiac Mechanical Simulations

0018-9294 (c) 2013 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission. Seehttp://www.ieee.org/publications_standards/publications/rights/index.html for more information.

This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI10.1109/TBME.2014.2373399, IEEE Transactions on Biomedical Engineering

IEEE TRANSACTIONS ON BIOMEDICAL ENGINEERING VOL. , NO. , MONTH 0000 7

Method Mean inflation pressure Worst-case pressure Typical EDP Extreme inflationDirect discretization 4.25 kPa 0.11 kPa 18/20 0/20... with stabilization 90.1 kPa 0.12 kPa 18/20 18/20

Isochoric/deviatoric discretization 95.1 kPa 1.3 kPa 20/20 19/20... with stabilization 95.2 kPa 3.8 kPa 20/20 19/20

Penalty method 85.2 kPa 0.01 kPa 18/20 17/20

TABLE IISUMMARY OF RESULTS FOR INFLATION OF BIVENTRICULAR MESHES. RESULTS SHOW KEY METRICS OF STABILITY FOR THE DIRECT

DISCRETIZATION COMPARED WITH OUR STABILIZED METHOD, DISCRETIZATION USING THE ISOCHORIC/DEVIATORIC SPLIT, AND A PENALTY METHOD.FOR EACH SETTING, WE SHOW THE NUMBER OF CASES THAT SUCCEED UP TO TYPICAL END-DIASTOLIC PRESSURE (1 KPA), AND THE NUMBER OF CASESTHAT REACH EXTREME INFLATION (100 KPA), AS WELL AS THE MEAN PRESSURE REACHED WHEN AVERAGING RESULTS OVER ALL 20 CASES, AND THEWORST CASE PRESSURE REACHED OVER ALL CASES. FOR THE STABILIZATION WE USE κ = 5PCAV , WHILE THE PENALTY METHOD USES κ = 1000 KPA.

Fig. 5. Influence of the penalty term on stability and computational cost. These plots show average performance over 20 cases for varying constantfactor κc and pressure dependent factor κp, where κ = κc + κp Pcav. The number in the middle of the squares shows the average number of Jacobiansrequired in successful simulations only. Letters in parenthesis indicate the meshes which failed to reach 1 kPa in these settings, and only include a,b,d as allother cases succeeded in all settings. Panel (a) shows these results for the direct discretization, and panel (b) for the deviatoric/isochoric formulation. Notethat the choice of 1 kPa pressure implies solutions to the deformation on top-left to bottom-right diagonals of identical κc + κp are completely identical,while total computational cost shown here can differ significantly.

shown to be effective for achieving high computational per-formance, shows its limitations in this challenging test case.Specifically, in isovolumetric relaxation, the decreasing pres-sure will result in a decreasing penalty, which we hypothesisedis the cause for their large number of failed simulations in IVR.This inspired a further test in which the parameter κ can notdecrease, resulting in a total of 95% of successful simulationsfor both activation patterns.

III. DISCUSSION

In this paper we have presented, to our knowledge, thefirst report of the performance of the different availablenumerical schemes on detailed simulations of the entire car-diac cycle. This analysis revealed the limited performance ofseveral existing numerical methods for the simulation of morechallenging heterogeneous cardiac simulations. In addition,we have proposed alternative ‘stabilized’ schemes which are

straightforward to implement and significantly increases thestability of simulations during extreme inflation, ejection, andisovolumetric relaxation in personalized cardiac geometries.

When analyzing the effects of the proposed methods us-ing an idealized cylinder inflation test, we showed that ourstabilized method improves the accuracy of solving the in-compressibility constraint while converging as expected fordeformation and volume measurements. In our test of biven-tricular mesh inflation, maximal inflation pressure was 6.6 kPaacross all meshes when applying the commonly used directdiscretization, When adding the stabilization term or usingthe deviatoric/isochoric formulation, at least 90% of the casesinflate to extreme pressures of 100 kPa. This is a significantimprovement over the previous results reported by Lamata etal. [10] using the same kind of cardiac meshes, and shows thecomplex interaction between mesh quality and the choice ofnumerical methods.

Page 8: Improving the Stability of Cardiac Mechanical Simulations

0018-9294 (c) 2013 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission. Seehttp://www.ieee.org/publications_standards/publications/rights/index.html for more information.

This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI10.1109/TBME.2014.2373399, IEEE Transactions on Biomedical Engineering

IEEE TRANSACTIONS ON BIOMEDICAL ENGINEERING VOL. , NO. , MONTH 0000 8

Homogeneous activationMethod Succeed Fail in inflation Fail in IVC Fail in ejection Fail in IVRDirect discretization 16 2 1 1... with stabilization 17 2 1... with stabilization, non-decreasing 17 2 1Isochoric/deviatoric discretization 5 1 11 3... with stabilization 16 1 3... with stabilization, non-decreasing 19 1

Septal activationMethod Succeed Fail in inflation Fail in IVC Fail in ejection Fail in IVRDirect discretization 0 2 1 11 6... with stabilization 1 2 1 16... with stabilization, non-decreasing 17 2 1Isochoric/deviatoric discretization 0 1 10 9... with stabilization 0 1 19... with stabilization, non-decreasing 19 1

TABLE IIIRESULTS FOR WHOLE CYCLE SIMULATIONS. THE LEFT-MOST COLUMN INDICATES THE NUMBER OF SIMULATIONS THAT COMPLETE SUCCESSFULLY

TO END ISOVOLUMETRIC RELAXATION. LETTERS IN THE OTHER COLUMNS INDICATE THE NUMBER OF MESHES FOR WHICH THE SIMULATION FAILED TOCONVERGE IN A CERTAIN CARDIAC PHASE, AND MATCH THOSE FROM FIGURE 5. FOR THE STABILIZATION WE USE κ = 5PCAV , AND FOR

‘NON-DECREASING’ ENTRIES WE DO NOT ALLOW κ TO DECREASE DURING THE SIMULATION.

Unlike its good performance in diastolic inflation, theisochoric/deviatoric formulation shows relatively poor per-formance in running whole-cycle models on personalizedpatient geometries. As our proposed stabilized method per-forms better on these simulations, and is competitive with theisochoric/deviatoric formulation for extreme inflation, it maybe a good choice with a lower implementation barrier for mod-ifying an existing finite element implementation to be morestable. However, the proposed stabilization can also be easilyadded to an implementation using the isochoric/deviatoricdiscretization, and this combination gives the best results inboth diastolic inflation tests, and completing whole cycles.

From the range of tests shown for the effects of the extraparameter κ we can derive some practical guidelines for thechoice of stabilization term. Although the stabilization termdoes not add significant computational cost by itself, a highpenalty term can in practice lead to slower convergence ofthe nonlinear solver. To limit this effect, we recommend usinga penalty with a constant factor of 1-5 kPa and a pressure-dependent factor of 1-25.

As mentioned previously, although the observation of neg-ative det F in the Newton iteration process was the maininspiration for the stabilized method, the connection betweenthis problem and solver stability remains to be fully elucidated.We identified the main cause of solver non-convergence to bea near-singular Jacobian leading to exploration of nonphysicaland self-intersecting trial solutions, and numerical overflowof the stress tensor. The contribution of the cubic orderdeformation to the incompressibility constraint is likely toregularize the deformation, and may help to prevent localextremes in stress and strain. Underlying these erratic failureof the Newton solver for more extreme deformations is theloss of ellipticity for the direct discretization, while moreconsistent failure at extreme deformations is correlated withloss of numerical precision due to high condition numbers.

In this study we focussed our attention on our preferredapplication of cardiac mechanics. Our focus on this particularapplication may have introduced a number of limitations.Firstly, we have used the Guccione constitutive law in our

test problems as it is relevant to our research and is similar tomost other constitutive laws used in cardiac mechanics [43],as they have similar exponential strain energy functions. If thenonconvergence is indeed related to a loss of ellipticity, dif-ferent types of constitutive laws may not require the stabilizedmethod. This makes our method more suitable for biologicalmaterials with complex anisotropic properties. Indeed, limitedtests using a Neo-Hookean law showed that very extreme andnon-physical deformations are required before the solver fails.However, as shown by Pathmanathan et al. [25], any strainenergy function will lose strong ellipticity when sufficientactive tension is added. The results shown in this paper whensolving a full cardiac cycle with the isochoric/deviatoric for-mulation, shows that although this method has been analyzedin detail for passive mechanics, the addition of active tensionintroduces unique challenges which have not been sufficientlyinvestigated. Another aspect of stability is the possibility forsnap-through solutions, in which one pressure yields multiplepossible solutions for the deformation. These issues can beaddressed by the use of arc-length methods [20]. However,for our constitutive laws such phenomena are not expectedbecause total strain energy grows exponentially in strain whileexternal work grows proportional to surface area. Indeed, oursolutions show no indication of sudden jumps in volume orsolver-independent points of failure.

In conclusion, in this study we have proposed a newformulation of incompressible large deformation mechanics.The method can improve the stability of simulations whilemaintaining computational efficiency.

REFERENCES

[1] M. P. Nash and A. V. Panfilov, “Electromechanical model of excitabletissue to study reentrant cardiac arrhythmias,” Prog. Biophys. Mol. Biol.,vol. 85, no. 2-3, pp. 501–522, 2004.

[2] M. Sermesant, K. Rhode, G. I. Sanchez-Ortiz, O. Camara, R. An-driantsimiavona, S. Hegde, D. Rueckert, P. Lambiase, C. Bucknall,E. Rosenthal, H. Delingette, D. L. G. Hill, N. Ayache, and R. Razavi,“Simulation of cardiac pathologies using an electromechanical biventric-ular model and XMR interventional imaging,” Med. Image Anal., vol. 9,no. 5, pp. 467–480, Oct. 2005.

Page 9: Improving the Stability of Cardiac Mechanical Simulations

0018-9294 (c) 2013 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission. Seehttp://www.ieee.org/publications_standards/publications/rights/index.html for more information.

This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI10.1109/TBME.2014.2373399, IEEE Transactions on Biomedical Engineering

IEEE TRANSACTIONS ON BIOMEDICAL ENGINEERING VOL. , NO. , MONTH 0000 9

[3] S. Land, S. A. Niederer, J. M. Aronsen, E. K. S. Espe, L. Zhang, W. E.Louch, I. Sjaastad, O. M. Sejersted, and N. P. Smith, “An analysis ofdeformation-dependent electromechanical coupling in the mouse heart,”J. Physiol., vol. 590, no. 18, pp. 4553–4569, Sep. 2012.

[4] F. Sheikh, K. Ouyang, S. G. Campbell, R. C. Lyon, J. Chuang,D. Fitzsimons, J. Tangney, C. G. Hidalgo, C. S. Chung, H. Cheng,N. D. Dalton, Y. Gu, H. Kasahara, M. Ghassemian, J. H. Omens, K. L.Peterson, H. L. Granzier, R. L. Moss, A. D. McCulloch, and J. Chen,“Mouse and computational models link mlc2v dephosphorylation toaltered myosin kinetics in early cardiac disease,” J. Clin. Invest., vol.122, no. 4, pp. 1209–1221, Apr. 2012.

[5] S. Land, W. E. Louch, S. A. Niederer, J. M. Aronsen, G. A.Christensen, I. Sjaastad, O. M. Sejersted, and N. P. Smith, “Beta-adrenergic stimulation maintains cardiac function in serca2 knockoutmice,” Biophys. J., vol. 104, no. 6, pp. 1349–1356, Mar. 2013.

[6] A. Krishnamurthy, C. T. Villongco, J. Chuang, L. R. Frank, V. Nigam,E. Belezzuoli, P. Stark, D. E. Krummen, S. Narayan, J. H. Omens,A. D. McCulloch, and R. C. Kerckhoffs, “Patient-specific models ofcardiac biomechanics,” J. Comput. Phys., vol. 244, pp. 4–21, Jul. 2013.

[7] T. P. Usyk and A. D. McCulloch, “Electromechanical model of cardiacresynchronization in the dilated failing heart with left bundle branchblock,” J. Electrocardiol., vol. 36, pp. 57–61, 2003.

[8] S. A. Niederer, G. Plank, P. Chinchapatnam, M. Ginks, P. Lamata, K. S.Rhode, C. A. Rinaldi, R. Razavi, and N. P. Smith, “Length-dependenttension in the failing heart and the efficacy of cardiac resynchronizationtherapy,” Cardiovascular Research, vol. 89, no. 2, p. 336, 2011.

[9] S. Niederer, P. Lamata, G. Plank, P. Chinchapatnam, M. Ginks,K. Rhode, C. Rinaldi, R. Razavi, and N. Smith, “Analyses of theredistribution of work following cardiac resynchronisation therapy ina patient specific model,” PLoS ONE, vol. 7, no. 8, 2012.

[10] P. Lamata, I. Roy, B. Blazevic, A. Crozier, S. Land, S. A. Niederer,D. R. Hose, and N. P. Smith, “Quality metrics for high order meshes:analysis of the mechanical simulation of the heart beat,” IEEE Trans.Med. Imag., vol. 32, no. 1, pp. 130–138, Jan. 2013.

[11] P. Lamata, S. Niederer, D. Nordsletten, D. C. Barber, I. Roy, D. R.Hose, and N. Smith, “An accurate, fast and robust method to generatepatient-specific cubic hermite meshes,” Med. Image Anal., vol. 15,no. 6, pp. 801–813, Dec. 2011.

[12] J. Xi, P. Lamata, S. Niederer, S. Land, W. Shi, X. Zhuang, S. Ourselin,S. G. Duckett, A. K. Shetty, C. A. Rinaldi, D. Rueckert, R. Razavi,and N. P. Smith, “The estimation of patient-specific cardiac diastolicfunctions from clinical measurements,” Med. Image Anal., vol. 17, no. 2,pp. 133–146, Feb. 2013.

[13] J. C. Simo, R. L. Taylor, and K. S. Pister, “Variational andprojection methods for the volume constraint in finite deformationelasto-plasticity,” Comput. Methods Appl. Mech. Eng., vol. 51, no. 1,pp. 177–208, 1985.

[14] J. C. Simo and R. L. Taylor, “Quasi-incompressible finite elasticityin principal stretches. continuum basis and numerical algorithms,”Comput. Methods Appl. Mech. Eng., vol. 85, no. 3, pp. 273–310, Feb.1991.

[15] C. Miehe, “Aspects of the formulation and finite element implementationof large strain isotropic elasticity,” Int. J. Numer. Methods Eng., vol. 37,no. 12, pp. 1981–2004, 1994.

[16] P. J. Flory, “Thermodynamic relations for high elastic materials,”Transactions of the Faraday Society, vol. 57, no. 0, pp. 829–838, Jan.1961.

[17] P. Wriggers, Nonlinear finite element methods. Springer, 2008.[18] S. Goktepe, S. N. S. Acharya, J. Wong, and E. Kuhl, “Computational

modeling of passive myocardium,” Int. J. Numer. Method. Biomed.Eng., vol. 27, no. 1, pp. 1–12, 2011.

[19] H. M. Wang, H. Gao, X. Y. Luo, C. Berry, B. E. Griffith, R. W.Ogden, and T. J. Wang, “Structure-based finite strain modelling of thehuman left ventricle in diastole,” Int. J. Numer. Method. Biomed. Eng.,vol. 29, no. 1, pp. 83–103, 2013.

[20] G. A. Holzapfel, Nonlinear solid mechanics: a continuum approach forengineering. Wiley-Blackwell, 2000.

[21] J. Bonet and R. D. Wood, Nonlinear Continuum Mechanics for FiniteElement Analysis, 2nd ed. Cambridge University Press, Mar. 2008.

[22] M. P. Nash and P. J. Hunter, “Computational mechanics of the heart,”Journal of elasticity, vol. 61, no. 1, pp. 113–141, 2000.

[23] T. P. Usyk, I. J. LeGrice, and A. D. McCulloch, “Computational modelof three-dimensional cardiac electromechanics,” Comput. Vis. Sci., vol. 4,no. 4, pp. 249–257, 2002.

[24] H. Watanabe, S. Sugiura, H. Kafuku, and T. Hisada, “Multiphysicssimulation of left ventricular filling dynamics using fluid-structure

interaction finite element method,” Biophys. J., vol. 87, no. 3, pp. 2074–2085, 2004.

[25] P. Pathmanathan, S. J. Chapman, D. J. Gavaghan, and J. P. Whiteley,“Cardiac electromechanics: the effect of contraction model on themathematical problem and accuracy of the numerical scheme,” Q. J.Mech. Appl. Math., vol. 63, no. 3, pp. 375–399, 2010.

[26] D. Nordsletten, S. Niederer, M. Nash, P. Hunter, and N. Smith,“Coupling multi-physics models to cardiac mechanics,” Prog. Biophys.Mol. Biol., vol. 104, no. 13, pp. 77–88, Jan. 2011.

[27] S. Land, S. A. Niederer, and N. P. Smith, “Efficient computationalmethods for strongly coupled cardiac electromechanics,” IEEE Trans.Biomed. Eng., vol. 59, no. 5, pp. 1219 –1228, May 2012.

[28] H. Xia, K. Wong, and X. Zhao, “A fully coupled model forelectromechanics of the heart,” Comput Math. Methods Med., vol.2012, Oct. 2012.

[29] R. C. P. Kerckhoffs, J. H. Omens, and A. D. McCulloch, “A singlestrain-based growth law predicts concentric and eccentric cardiacgrowth during pressure and volume overload,” Mechanics ResearchCommunications, vol. 42, pp. 40–50, Jun. 2012.

[30] H. Dal, S. Goktepe, M. Kaliske, and E. Kuhl, “A fully implicit finiteelement method for bidomain models of cardiac electromechanics,”Comput. Methods Appl. Mech. Eng., vol. 253, pp. 323–336, Jan. 2013.

[31] A. Cheng, F. Langer, F. Rodriguez, J. C. Criscione, G. T. Daughters,D. C. Miller, and N. B. Ingels Jr, “Transmural cardiac strains in thelateral wall of the ovine left ventricle,” Amer. J. Physiol. Heart Circ.Physiol., vol. 288, no. 4, pp. H1546–1556, Apr. 2005.

[32] J. M. Guccione, A. D. McCulloch, and L. K. Waldman, “Passivematerial properties of intact ventricular myocardium determined from acylindrical model,” J. Biomech. Eng., vol. 113, no. 1, pp. 42–55, Feb.1991.

[33] I. J. Legrice, P. J. Hunter, and B. H. Smaill, “Laminar structure of theheart: a mathematical model,” Amer. J. Physiol. Heart Circ. Physiol.,vol. 272, no. 5, pp. H2466–2476, May 1997.

[34] J. A. Weiss, B. N. Maker, and S. Govindjee, “Finiteelement implementation of incompressible, transversely isotropichyperelasticity,” Comput. Methods Appl. Mech. Eng., vol. 135, no. 12,pp. 107–128, Aug. 1996.

[35] J. Helfenstein, M. Jabareen, E. Mazza, and S. Govindjee, “On non-physical response in models for fiber-reinforced hyperelastic materials,”Int. J. Solids Struct., vol. 47, no. 16, pp. 2056–2061, Aug. 2010.

[36] S. Federico, A. Grillo, G. Giaquinta, and W. Herzog, “Convexfung-type potentials for biological tissues,” Meccanica, vol. 43, no. 3,pp. 279–288, Jun. 2008.

[37] G. A. Holzapfel and R. W. Ogden, “Constitutive modelling of passivemyocardium: a structurally based framework for material characteriza-tion,” Phil. Trans. R. Soc. A, vol. 367, no. 1902, p. 3445, 2009.

[38] P. Lamata, M. Sinclair, E. Kerfoot, A. Lee, A. Crozier, B. Blazevic,S. Land, A. J. Lewandowski, D. Barber, S. Niederer, and N. Smith, “Anautomatic service for the personalization of ventricular cardiac meshes,”J. R. Soc. Interface, vol. 11, no. 91, p. 20131023, Feb. 2014.

[39] E. Kerfoot, P. Lamata, S. Niederer, R. Hose, J. Spaan, and N. Smith,“Share and enjoy: anatomical models database - generating and sharingcardiovascular model data using web services,” Medical & BiologicalEngineering & Computing, pp. 1–10, 2013.

[40] D. A. Kass, M. Midei, J. Brinker, and W. L. Maughan, “Influenceof coronary occlusion during PTCA on end-systolic and end-diastolicpressure-volume relations in humans.” Circulation, vol. 81, no. 2, pp.447–460, Jan. 1990.

[41] N. Westerhof and G. Elzinga, “Normalized input impedance and arterialdecay time over heart period are independent of animal size,” Am. J.Physiol. Regul. Integr. Comp. Physiol., vol. 261, no. 1, pp. R126–133,Jul. 1991.

[42] R. C. P. Kerckhoffs, P. H. M. Bovendeerd, F. W. Prinzen, K. Smits, andT. Arts, “Intra- and interventricular asynchrony of electromechanicsin the ventricularly paced heart,” J Eng. Math., vol. 47, no. 3/4, pp.201–216, Dec. 2003.

[43] H. Schmid, M. P. Nash, A. A. Young, and P. J. Hunter, “Myocardialmaterial parameter estimationa comparative study for simple shear,” J.Biomech. Eng., vol. 128, p. 742, 2006.