Вы находитесь на странице: 1из 13

Accepted Manuscript

The Living Heart Project: A robust and integrative simulator for human heart function

Brian Baillargeon, Nuno Rebelo, David D. Fox, Robert L. Taylor, Ellen Kuhl

PII: S0997-7538(14)00056-4
DOI: 10.1016/j.euromechsol.2014.04.001
Reference: EJMSOL 3055

To appear in: European Journal of Mechanics / A Solids

Received Date: 22 February 2014


Revised Date: 7 April 2014
Accepted Date: 7 April 2014

Please cite this article as: Baillargeon, B., Rebelo, N., Fox, D.D., Taylor, R.L., Kuhl, E., The Living Heart
Project: A robust and integrative simulator for human heart function, European Journal of Mechanics / A
Solids (2014), doi: 10.1016/j.euromechsol.2014.04.001.

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to
our customers we are providing this early version of the manuscript. The manuscript will undergo
copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please
note that during the production process errors may be discovered which could affect the content, and all
legal disclaimers that apply to the journal pertain.
Graphical Abstract (for review)
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE
EP
C
AC
*Highlights (for review)
ACCEPTED MANUSCRIPT

HIGHLIGHTS

The Living Heart Project:


A robust and integrative simulator for human heart function

PT
Brian Baillargeon, Nuno Rebelo, David D. Fox, Robert L. Taylor, Ellen Kuhl

RI
► We created a human heart model with four chambers and four valves
► Filling of the chambers is controlled by the opening and closing of the valves

SC
► Ejection is controlled by the interaction of electrical and mechanical fields
► Our model predicts pressure-volume loops, which agree nicely with clinical
observations

U
► Our model opens opportunities to virtually probe parameters and guide treatment
planning
AN
M
D
TE
C EP
AC
*Manuscript with Highlighted Changes
Click here to view linked References

ACCEPTED MANUSCRIPT

The Living Heart Project:


A robust and integrative simulator for human heart function

Brian Baillargeona , Nuno Rebeloa , David D. Foxb , Robert L. Taylorc , Ellen Kuhld
a DassaultSystèmes Simulia Corporation, Fremont, CA 94538, USA
b DassaultSystèmes Simulia Corporation, Providence, RI 02909, USA
c Department of Civil and Environmental Engineering, University of California at Berkeley, Berkeley, CA 94720, USA
d Departments of Mechanical Engineering, Bioengineering, and Cardiothoracic Surgery, Stanford University, Stanford, CA 94305, USA

PT
RI
Abstract
The heart is not only our most vital, but also our most complex organ: Precisely controlled by the interplay of electrical and
mechanical fields, it consists of four chambers and four valves, which act in concert to regulate its filling, ejection, and overall pump

SC
function. While numerous computational models exist to study either the electrical or the mechanical response of its individual
chambers, the integrative electro-mechanical response of the whole heart remains poorly understood. Here we present a proof-
of-concept simulator for a four-chamber human heart model created from computer topography and magnetic resonance images.
We illustrate the governing equations of excitation-contraction coupling and discretize them using a single, unified finite element

U
environment. To illustrate the basic features of our model, we visualize the electrical potential and the mechanical deformation
across the human heart throughout its cardiac cycle. To compare our simulation against common metrics of cardiac function, we
AN
extract the pressure-volume relationship and show that it agrees well with clinical observations. Our prototype model allows us to
explore and understand the key features, physics, and technologies to create an integrative, predictive model of the living human
heart. Ultimately, our simulator will open opportunities to probe landscapes of clinical parameters, and guide device design and
treatment planning in cardiac diseases such as stenosis, regurgitation, or prolapse of the aortic, pulmonary, tricuspid, or mitral valve.
M

Keywords: Electro-mechanics; Excitation-contraction; Cardiac mechanics; Finite element analysis; Abaqus


D

1. Motivation

The human heart beats about 100,000 times daily, 30 million


TE

times annually, and 2.5 billion times in an average lifetime [5].


Only marginally larger than a fist, it is capable of pumping superior ! aortic !
vena cava! arch!
7,000 liters of blood per day, 2.5 million per year, and 200 mil-
lion throughout an individual’s life span [31]. Figure 1 shows
EP

an anatomic model of the human heart created from computer pulmonary!


tomography and magnetic resonance images [50]. Our heart artery!
consists of four chambers, the left and right atria and the left
and right ventricles, connected by four valves. Figure 2 illus-
right ! left!
C

trates the four valves, the tricuspid and mitral valves, which
atrium! atrium!
connect the right and left atria to the right and left ventricles,
AC

and the pulmonary and aortic valves, which connect the right
and left ventricles to the pulmonary and systemic circulation
[50]. The coordinated opening and closing of these valves regu-
lates the filling of the chambers, while the interplay of electrical right ! left!
and mechanical fields controls their proper ejection. Disturbed ventricle! ventricle!
valvular opening, stenosis, disturbed closing, regurgitation, dis-
turbed electrical signals, arrhythmias, and reduced mechanical
function, heart failure, can have devastating physiological con-
sequences [7]. Modeling the interplay between electrical ex- Figure 1: Anatomic model of the human heart created from computer tomog-
raphy and magnetic resonance images. The model displays the characteristic
citation and mechanical contraction provides insight into these anatomic features: The aortic arch, the pulmonary artery, and the superior vena
complex phenomena and holds the potential to improve treat- cava; the two upper chambers, the left and right atria; and the two lower cham-
ment for the millions of people affected by cardiac disease [24]. bers, the left and right ventricles; adopted with permission from [50].

Preprint submitted to European Journal of Mechanics A/Solids April 6, 2014


ACCEPTED MANUSCRIPT

weak forms of the governing equations, their temporal and spa-


tial discretizations, their linearizations, and the handling of their
internal variables. In Section 4, we document the creation of
aortic ! pulmonary! our human heart model including the solid model, the finite
arch! artery! element model, the muscle fiber model, the fluid model, and a
summary of all model parameters. In Section 5 we illustrate the
pulmonary! simulation of an entire cardiac cycle. We close with a discus-
aortic! valve! sion of the results, the limitations, an outlook, and some con-
valve! clusions in Section 6.
tricuspid! mitral!

PT
valve! valve!
2. Continuum model

right ! left! We illustrate the continuum model of electro-mechanical cou-

RI
ventricle! ventricle! pling by briefly summarizing the kinematic equations, the bal-
ance equations, and the constitutive equations of excitation-
contraction coupling.

SC
Figure 2: Circulatory model of the human heart created from computer tomog- 2.1. Kinematic equations
raphy and magnetic resonance images. The model displays the characteristic
circulatory features: The tricuspid and mitral valves, which connect the right To characterize the kinematics of finite deformation, we intro-
and left atria to the right and left ventricles; and the pulmonary and aortic duce the deformation map ϕ, which maps particles X from the

U
valves, which connect the right and left ventricles to the pulmonary and sys- undeformed reference configuration to particles x = ϕ ( X, t ) in
temic circulation; adopted with permission from [50].
the deformed configuration [22]. Its derivative with respect to
AN
the undeformed coordinates X defines the deformation gradi-
ent,
Heart disease is the primary cause of death in the industrialized F = ∇ϕ , (1)
nations, claiming more than 16 million lives worldwide every
M

year [2]. The origin of the heart disease is often local: Fibril- along with its Jacobian, J = det(F). We can then introduce the
lation and myocardial infarction are classical examples of local right Cauchy-Green deformation tensor,
electrical and mechanical dysfunction. However, irrespective
of its nature and initial location, cardiac disease almost always C = Ft · F, (2)
D

progresses to affect the entire heart, and eventually impairs the


electrical and mechanical function of all four chambers [31]. and four of its invariants, which result from its projection onto
the unit tensor I, and the undeformed myocardial fiber and sheet
TE

To understand the fundamental pathologies of different forms


of cardiac disease and optimize their treatment options, it is unit directions f 0 and s0 ,
critical to model the entire heart as a whole, rather than study-
ing the diseased subsystem in complete isolation [44]. II = C : I Iff = C : [ f 0 ⊗ f 0 ]
(3)
Iss = C : [s0 ⊗ s0 ] Ifs = C : [ f 0 ⊗ s0 ] .
EP

To date, numerous computational models exist to simulate the


behavior of the left ventricle [11, 28, 32, 41], fewer model exist
The invariants Iff and Iss take the interpretation of the fiber and
to simulate both the ventricles [4, 26, 29, 45], and only very few
sheet stretch squared as the squared lengths of the deformed
models exist to simulate the human heart with all four chambers
fiber and sheet vectors, f = F· f 0 and s = F·s0 , and Ifs indicates
C

[20, 44]. Creating whole heart models remains challenging be-


the fiber-sheet shear. In what follows, we denote the material
cause the atrial wall is about an order of magnitude thinner than
time derivative as {˙◦} = d{◦}/dt and the material gradient and
the ventricular wall. In addition, the atria are typically entan-
AC

divergence as ∇{◦} = d{◦}/dX and Div {◦} = d{◦}/dX : I.


gled and their geometry can be quite complex [20]. This not
only complicates image segmentation, but also atrial discretiza-
tion and meshing. Here we create a finite element model of the 2.2. Balance equations
whole heart on the basis of existing anatomic and circulatory We characterize the electrical problem through the mono-
models illustrated in Figures 1 and 2. This allows us to model domain version of the FitzHugh-Nagumo equations [12, 35] for
all four chambers as electrically excitable, deformable, hypere- the electrical potential φ and the mechanical problem through
lastic, electroactive bodies connected via in- and out-flow con- the balance of linear momentum for the deformation ϕ,
ditions of viscous resistance type.
The remainder of this manuscript is organized as follows: In φ̇ = Div ( q ) + f φ
(4)
Section 2, we summarize the continuum model of electro- 0 = Div ( P ) + f ϕ .
mechanical coupling based on the kinematic equations, the bal-
ance equations, and the constitutive equations. In Section 3, Here, q is the electrical flux, f φ is the transmembrane current,
we illustrate our computational model, based on the strong and P is the Piola stress, and f ϕ is the external mechanical force.
2
ACCEPTED MANUSCRIPT

2.3. Constitutive equations 3. Computational model


To close the set of equations, we specify the constitutive equa- In this section, we illustrate the finite element discretization of
tions for the electrical flux q, the transmembrane current f φ , the governing equations, demonstrate their consistent lineariza-
the Piola stress P, and the external mechanical force f ϕ . We tion, and discuss the handling of their internal variables [6].
introduce the electrical flux proportional to the gradient of the
electrical field, 3.1. Strong and weak forms
q = D · ∇φ , (5) To derive the weak form of the governing equations, we refor-
where D = diso I + dani f 0 ⊗ f 0 denotes the conductivity ten- mulate the electrical and mechanical balance equations (4) in
sor, which consists of an isotropic contribution diso and an their residual forms and introduce the electrical and mechanical

PT
anisotropic contribution dani to account for faster conductivity residuals Rφ and Rϕ throughout the entire cardiac domain B0 .
along the fiber direction f 0 [9]. For the transmembrane current, .
Rφ = φ̇ − Div (q) − f φ = 0
. (11)
f φ = c φ [φ − α][φ − 1] − rφ , (6) Rϕ = − Div(P) − f ϕ = 0 .

RI
We prescribe Dirichlet boundary conditions φ = φ̄ and ϕ = ϕ̄
we assume a cubic polynomial, c φ [φ−α][φ−1], which controls
on the Dirichlet boundary and Neumann boundary conditions
the fast upstroke of the action potential through the parameters c
q · N = t¯φ and P · N = t̄ ϕ on the Neumann boundary with

SC
and α [12, 35], and of a coupling term, which controls the slow
outward normal N. For simplicity, we assume that all Neumann
repolarization through the recovery variable r [3]. We treat the
boundary conditions are homogeneous, t¯φ = 0 and t̄ ϕ = 0. We
recovery variable as internal variable, which evolves according
multiply the residuals (11) by the scalar- and vector-valued test
to the following equation,
functions, δφ and δϕ, integrate them over the domain B0 , and

U
ṙ = [ γ + r µ1 /[µ2 + φ] ] [−r − c φ [φ − β − 1]] , (7) integrate the flux terms by parts to obtain the weak forms of the
electrical and mechanical problems,
AN
where the recovery parameters γ, µ1 , µ2 and β control the resti-
.
Z Z Z
tution behavior [3]. We assume that the tissue stress consists of Gφ = δφ φ̇ dV + ∇δφ · q dV − δφ f φ dV = 0
passive and active contributions, B0 ZB0 ZB0 (12)
.
Gϕ = ∇δϕ : P dV − δϕ · f ϕ dV = 0
M

P = Ppas + Pact . (8) B0 B0

Next, we discretize the weak forms (12) in time and space.


For the passive stress Ppas , we postulate a Holzapfel-type re-
sponse along the fiber and sheet directions f 0 and s0 [23]. For
D

3.2. Temporal and spatial discretizations


the active stress Pact , we assume that the electrical field initi-
To discretize the weak form of the electrical problem (12.1)
ates an active muscle traction T act , which generates contraction
in time, we partition the time interval of interest T into nstep
TE

along the fiber and sheet directions [47].


discrete subintervals [tn , tn+1 ] of length ∆t = tn+1 − tn ,
Ppas = κ [ J 2 − 1 ] F−t nstep
+ a exp (b[II − 3]) F T = U [tn , tn+1 ], (13)
n=1
+ 2 aff [Iff − 1] exp (bff [Iff − 1]2 ) f ⊗ f 0
EP

(9) and adopt a finite difference discretization in combination with


+ 2 ass [Iss − 1] exp (bss [Iss − 1]2 ) s ⊗ s0
a classical implicit Euler backward scheme to determine the
+ afs Ifs exp (bfs Ifs2 ) [ f ⊗ s0 + s ⊗ f 0 ]
electrical potential φ at the current time point tn+1 ,
Pact = T act [ νff f ⊗ f 0 + νss s ⊗ s0 ] .
C

φ̇ = [ φ − φn ] / ∆t . (14)
Here, κ is the bulk modulus, a, b, aff , bff , ass , bss , afs , bfs are
AC

the parameters of the orthotropic Holzapfel model [18, 23], and To discretize the weak forms of the electrical and mechanical
νff and νss are the weighting factors for active stress generation problems (12.1) and (12.2) in space, we partition the domain of
[40, 47]. The active muscle traction is driven by changes in the interest B0 into nel discrete subdomains Be0 ,
electrical potential and obeys the following evolution equation nel
[16], B0 = U Be0 (15)
e=1
Ṫ act = (φ) [kT [φ − φr ] − T act ] . (10)
and adopt a finite element discretization in combination with a
The parameters kT and φr control the maximum active force
classical Bubnov-Galerkin scheme to discretize the test func-
and the resting potential [36]. The activation function  = 0 +
tions δφ and δϕ and trial functions φ and ϕ in space [16],
[∞ − 0 ] exp(− exp(−ξ[φ − φ̄])) ensures a smooth activation of
the muscle traction T act in terms of the limiting values 0 at δφ = i=1 Ni δφi
P
δϕ = j=1 N j δϕ j
P
φ → −∞ and and ∞ at at φ → +∞, the phase shift φ̄, and the (16)
φ = k=1 Nk φk ϕ = l=1 Nl ϕl .
P P
transition slope ξ [16]. In the following, we assume that we can
neglect the effects of external forces, f ϕ = 0. Here, N are the standard isoparametric shape functions.
3
ACCEPTED MANUSCRIPT

aortic ! aortic !
arch! arch!

superior ! pulmonary! superior ! pulmonary!


vena cava! artery! vena cava! artery!

right ! left! right ! left!


atrium! atrium! atrium! atrium!

PT
right ! left! right ! left!
ventricle! ventricle! ventricle! ventricle!

RI
SC
Figure 3: Solid model of the human heart with anatomic details including the Figure 4: Finite element model of the human heart discretized with 208,561
aortic arch, pulmonary artery and superior vena cava, left and right atria, and linear tetrahedral elements, 47,323 nodes, and 189,292 degrees of freedom, of
left and right ventricles. which 47,323 are electrical and 141,969 are mechanical.

U
3.3. Residuals and consistent linearization remains to specify the fluxes q and P and sources f φ and f ϕ
With the discretizations in time (14) and space (16), we can re- for the residuals (17.1) and (17.2) and their sensitivities with
AN
formulate the weak forms (12) as the discrete algorithmic resid- respect to the primary unknowns φ and ϕ for the iteration ma-
uals of the electrical and mechanical problems, trices (19.1) to (19.4) [10, 16].
φ
nel Z
φ − φn .
+ ∇Ni · q − Ni f φ dVe = 0
M

RI = A Ni 3.4. Internal variables


e=1 Be ∆ t
0
(17) To integrate the evolution equations of the recovery variable r
ϕ
nel Z
.
RJ = A ∇N j · P − N j f ϕ dVe = 0 and the active muscle traction T act in time, we treat both as
e=1 Be0
internal variables and update and store them locally on the in-
D

The operator A symbolizes the assembly of all element resid- tegration point level [16, 30]. To solve the nonlinear evolu-
uals at the element nodes i and j to the global residuals at the tion equation (7) for the recovery variable r, we locally adopt
TE

global nodes I and J. To solve for the unknown nodal electri- a finite difference discretization in combination with a classical
cal potential φI and mechanical deformation ϕ J , we could, for implicit Euler backward scheme [15],
example, adapt an incremental interative Newton-Raphson so-
lution strategy based on the consistent linearization of the gov- ṙ = [ r − rk ] / ∆t , (20)
EP

erning equations,
and introduce the local residual Rr ,
φ .
Kφφ Kφϕ
X X
RI + IK dφK + IL · dϕL = 0 µ1 r
Rr = r − rk − γ∆t + [r + c φ [φ − β − 1]]∆t = 0, (21)
K L
. (18) µ2 + φ
C

ϕ
Kϕφ Kϕϕ
X X
RJ + JK dφ K + JL · dϕL = 0 .
K L and its algorithmic linearization Kr ,
AC

The solution of this system of equations (18) with the discrete µ1


residuals (17) and the iteration matrices, Kr = 1 + [ 2 r − c φ[φ − β − 1] ]∆t . (22)
µ2 + φ
1
nel
Z " #
φφ
KIK = A Ni − dφ f φ Nk + ∇Ni · D · ∇Nk dVe to iteratively update the recovery variable as r ← r − Rr / Kr
e=1 Be
nel
Z 0 ∆t [29]. To solve the linear evolution equation (10) for the ac-
φϕ
KIL = A ∇Ni · d F q · ∇Nl dVe tive muscle traction T act , we again adopt a finite difference dis-
e=1 Be
nel
Z 0 (19) cretization in time together with an implicit Euler backward
ϕφ scheme,
K JK = A ∇N j · dφ Pact Nk dVe
e=1 Be
Z 0 Ṫ act = [ T act − T kact ] / ∆t (23)
nel
ϕϕ
K JL = A ∇N j · d F P · ∇Nl dVe , and solve the resulting equation directly to calculate the active
e=1 Be0
muscle traction at the current point in time,
defines the iterative update of the global vector of electrical and
mechanical unknowns φI ← φI + dφI and ϕ J ← ϕ J + dϕ J . It T act = [ T kact +  kT [ φ − φr ] ∆t ] / [ 1 +  ∆t ] , (24)
4
ACCEPTED MANUSCRIPT

aortic ! superior! right !


arch! vena cava! left !
atrium! atrium!

mitral valve!
left ! right! tricuspid valve!
atrium! atrium!

PT
left ! right! systemic! pulmonary
ventricle! ventricle! circulation! circulation!

right ! left !

RI
ventricle! ventricle!
aortic valve!
Figure 5: Muscle fiber model of the human heart with 208,561 discrete fiber

SC
and sheet directions interpolated and assigned to each integration point. Figure 6: Blood flow model of the human heart with surface-based fluid cavity
representation of the right atrium, right ventricle, left atrium, and left ventricle
connected through viscous resistance models of Windkessel type for the tri-
where the  = 0 + [∞ − 0 ] exp(− exp(−ξ[φ − φ̄])) [10]. Once cuspid valve, pulmonary circulation, mitral valve, aortic valve, and systemic
circulation.
we have determined the recovery variable r and the active mus-

U
cle traction T act , we calculate the electrical flux q from equa-
tion (5), the electrical source f φ from equation (6), the ac- at the endocardium, the inner wall, they point clockwise-
AN
tive stress Pact from equation (9), and the total stress P from downward. We interpolate the fiber and sheet directions from
equation (8) to evaluate the electrical and mechanical residuals generic fiber orientation illustrations and assign their discrete
(17). Last, we calculate the sensitivities dφ f φ [29], d F q = 0, values to each integration point of the finite element model [49].
dφ Pact = ∂φ T act [νff f ⊗ f 0 + νss s ⊗ s0 ], and d F P [18] for the
M

electrical and mechanical iteration matrices (19). 4.2. Fluid model


Figure 6 illustrates the blood flow model with a surface-based
4. Human heart model fluid cavity representation of the four chambers, the right
D

atrium, the right ventricle, the left atrium, and the left ventri-
In this section, we illustrate the creation of a solid model, cle [1]. These chambers are connected through five viscous
a finite element model, and a muscle fiber model from the resistance models of Windkessel type [4] representing the tri-
TE

anatomic model in Figure 1 and the creation of a fluid model cuspid valve, the pulmonary circulation, the mitral valve, the
from the circulatory model in Figure 2. This work was per- aortic valve, and the systemic circulation [43]. We adapt the
formed as part of the Living Heart Project. positions of these valves and the corresponding chamber vol-
umes from the detailed circulatory model in Figure 2 [50]. We
EP

4.1. Solid model neglect inertia effects and assume that the flow rate between
two neighboring chambers, qc→c+1 , is proportional to the pres-
Figure 3 shows the solid model of a human heart with well-
sure difference in the two chambers, pc − pc+1 , scaled by the
defined anatomic details including the aortic arch, the pul-
resistance Rc→c+1 [13],
C

monary artery and superior vena cava, the left and right atria,
and the left and right ventricles. We adapt the underlying geom- pc − pc+1
qc→c+1 = .
etry from the three-dimensional computer-aided design model
AC

Rc→c+1
in Figure 1 [50]. The model accurately defines the key features
To represent the four chambers, we create fluid cavities from
for our finite element analysis including detailed chamber vol-
cubes of unit volume and add the corresponding volume Vc [1]
umes and wall thicknesses.
§11.5. We then define the change in chamber volume,
Figure 4 illustrates the finite element model of the heart dis-
cretized with 208,561 linear tetrahedral elements and 47,323 V̇c = qc−1→c − qc→c+1 ,
nodes. This discretization introduces 47,323 electrical degrees
of freedom for the scalar-valued potential φ and 141,969 me- as the difference between influx qc−1→c and outflux qc→c+1 of
chanical degrees of freedom for the vector-valued deforma- the corresponding chamber. This simplified approach provides
tion ϕ resulting in 189,292 degrees of freedom in total. a natural coupling between the structural deformation and the
Figure 5 illustrates the corresponding muscle fiber model with fluid pressure, in which the fluid is represented exclusively
208,561 discrete fiber and sheet directions f 0 and s0 . The mus- in terms of the temporally varying blood pressure in the four
cle fibers wrap helically around the heart. At the epicardium, chambers. Spatial pressure variations or shear stresses cannot
the outer wall, muscle fibers point clockwise-upwards whereas be modeled with this fluid cavity representation.
5
ACCEPTED MANUSCRIPT

4.3. Solid and fluid model parameters 5.1. Electrical and mechanical fields
Table 1 summarizes the electrical, mechanical, electro-mecha- Figure 7 displays the spatio-temporal evolution of the electrical
nical, and flow parameters of the human heart simulation. potential, φ, the mechanical displacement, u = || ϕ− X ||, and the
muscle fiber strain, Eff = E : [ f 0 ⊗ f 0 ], across the human heart.
Table 1: Model parameters of the human heart simulation. The displacement u illustrates the magnitude of the displace-
ment vector u as the difference between the current position ϕ
electrical and initial position X. The muscle fiber strain Eff = Iff − 1 indi-
conduction diso = 2 mm2 /ms dani = 6 mm2 /ms [16] cates the strain along the muscle fiber direction f 0 . Initially the
excitation α = 0.01 γ = 0.002 [17] heart is at rest and its cells are negatively charged with a poten-
tial of φ = −80mV, see Figure 7, top row. The heart is excited

PT
β = 0.15 c =8 [29]
µ1 = 0.2 µ2 = 0.3 [3] from the sinoatrial node, a region between the right atrium and
the superior vena cava, which is the first region to depolarize
mechanical
with a potential of φ = +20mV. The excitation wave spreads
passive κ = 1, 000 kPa

RI
rapidly across the atria, arrests briefly at the atrialventricular
a = 0.496 kPa b = 7.209 [18] node, and continues to travel along the septum to rapidly excite
aff = 15.193 kPa bff = 20.417 [18] the left and right ventricles. After a short period of complete
ass = 3.283 kPa bss = 11.176 [18] depolarization, repolarization spreads gradually across the left

SC
afs = 0.662 kPa bfs = 9.466 [18] and right ventricles and atria to bring the heart back to its unex-
active νff = 1.0 νss = 0.4 [47] cited baseline state.
kT = 0.49 kPa/mV φr = −80 mV [10] The mechanical deformation clearly follows the electrical sig-
nal and spreads from the region that is excited first, the sinoa-

U
coupling
trial node, across the entire heart, see Figure 7, middle row.
activation 0 = 0.1/mV ∞ = 1.0 /mV [16] Once the heart is fully excited, the electrical potential is ho-
AN
ξ = 1/mV φ̄ = 0 mV [16] mogeneous across the heart; yet the mechanical deformation is
not. This clearly indicates the importance of the spatially vary-
blood flow
ing fiber orientation, which causes a local interplay between
blood ρ = 1.025·10−6 kg/mm3 [42]
cellular contraction and secondary effects induced by neighbor-
Rtv = 0.0010 kPa ms/mm3
M

valves [43] ing heart muscle fibers. As the electrical potential returns to its
Rmv = 0.0061 kPa ms/mm3 [43] baseline state, the deformation gradually decays, the heart re-
Rav = 0.0027 kPa ms/mm3 [43] laxes, and prepares for the next filling phase.
circulation Rpc = 0.0104 kPa ms/mm3 The muscle fiber strain mimics the effects of the overall defor-
D

Rsc = 0.0850 kPa ms/mm3 mation, projected onto the local fiber direction, see Figure 7,
bottom row. During systole, the muscle fibers contract rapidly
TE

and shorted up to 20% to induce ventricular ejection. This


For the electrical problem, we choose the initial conditions to moves the apex upward toward the base and induces a twist
φ = −80mV throughout the entire heart, except for a short ex- along the heart’s long axis. During diastole, the muscle fibers
citation phase, during which we locally increase the voltage be- gradually relax to allow for ventricular filling and the apex grad-
EP

yond the excitation threshold in the region of the sinoatrial node ually returns to its initial position.
located between the right atrium and the superior vena cava.
For the mechanical problem, we adopt a combination of dif- 5.2. Electrical potential throughout a cardiac cycle
ferent boundary conditions. At the epicardium, the outer wall,
Figure 8 illustrates the local temporal evolution of the electrical
C

we fix the heart in space through connector elements with a


potential φ for an individual cardiac muscle cell. During ex-
moderate elastic stiffness situated at the intersection between
citation, the cell depolarizes rapidly and its electrical potential
AC

the four-chamber heart and its vasculature [1] §31.2, similar to


increases from -80mV to +20mV within the order of millisec-
supporting the heart by linear elastic springs [16]. At the endo-
onds. This excitation causes the cell to contract. Unlike nerve
cardium, the inner wall, we control the pressure-volume rela-
cells, cardiac cells briefly plateau at the excited state before they
tion through surface-based fluid cavities [1] §11.5. We choose
begin to relax. During relaxation, cardiac cells gradually repo-
the initial conditions according to a preload step, during which
larize and return to their stable baseline state at -80mV.
we pressurize the right atrium and ventricle with 0.266kPa and
the left atrium and ventricle with 0.533kPa.
5.3. Mechanical deformation throughout a cardiac cycle
Figure 9 illustrates the dynamics of the heart’s long axis. Dur-
5. Results
ing ventricular ejection, the distance between apex and base
To demonstrate the ability of the proposed model to accurately decreases rapidly and the ventricles shorten by approximately
represent the basic features of cardiac excitation and contrac- 7mm. Shortening plateaus towards end systole to ensure that
tion, we simulate the electro-mechanical response of the human enough blood is ejected. During ventricular filling, the long
heart throughout a representative cardiac cycle. axis gradually returns to its initial length as the heart muscle
6
ACCEPTED MANUSCRIPT

electrical potential! -80mV! 20mV!

PT
RI
SC
mechanical displacement! 0mm! 10mm!

U
AN
M
D
TE

muscle fiber strain! -0.20! 0.20!


C EP
AC

Figure7:7:Spatio-temporal
Figure Spatio-temporalevolution
evolutionofofelectrical
electricalpotential,
potential, mechanical
mechanical displacement,
deformation, and andmuscle
musclefiberfiberstrain
strainacross
acrossthe
thehuman
humanheart.
heart.During
Duringsystole,
systole,thetheheart
heart
depolarizesrapidly
depolarizes rapidlyfrom
from-80mV
-80mVtoto+20mV,
+20mV,the themuscle
muscle fibers
fibers contract
contract and
and shorten
shorten up
up toto 20%
20% to to induce
induce ventricular
ventricularejection.
ejection.During
Duringdiastole,
diastole,the
theheart
heartrepolarizes
repolarizes
graduallyfrom
gradually from+20mV
+20mVtoto-80mV,
-80mV,the
themuscle
musclefibers
fibersrelax
relax and
and relengthen
relengthen to
to their
their initial
initial length
length to
to induce
induceventricular
ventricularfilling.
filling.

7
7
ACCEPTED MANUSCRIPT

20! relaxes.
Figure 10 illustrates the pressure-volume loop of the human
0! heart, the temporal evolution of the left ventricular pressure
gradual! and volume throughout a cardiac cycle. The first phase, ven-
electric potential [mV]!

repolarization! tricular filling, begins with the opening of the mitral valve and
-20!
continues towards end diastole, the point of maximum volume
and minimum pressure in the bottom right corner. The sec-
-40! ond phase, isovolumetric contraction, is characterized through
rapid a steep increase in pressure while the ventricular volume re-
depolarization! mains unchanged. Once the ventricular pressure exceeds the

PT
-60!
aortic pressure the aortic valve opens. This is the beginning of
the third phase, ventricular ejection, during which the volume
-80! of the ventricle decreases, while the pressure still remains high.
0! 100! 200! 300! 400! 500! 600! Ejection continues towards end systole, the point of minimum

RI
time [ms]!
volume and maximum pressure in the top left corner. The fourth
Figure 8: Temporal evolution of electrical potential. During excitation, car- phase, isovolumetric relaxation, begins with closure of the aor-
diac cells rapidly depolarize and the electrical potential increases from -80mV tic valve followed by a drastic pressure drop. A new cycle starts

SC
to +20mV within the order of milliseconds. During relaxation, cardiac cells with the reopening of the mitral valve and the beginning of ven-
gradually repolarize and return to the stable baseline state at -80mV.
tricular filling.
-6.0!
6. Discussion

U
long-axis shortening [mm]!

-4.0!
Modeling the human heart with all four chambers and all four
ventricular!
AN
filling! valves is increasingly recognized as a critical step towards re-
-2.0!
liable, predictive modeling of cardiac function [44]. Although
the origin of cardiac disease is often strictly local, the conse-
ventricular!
ejection! quences are typically spatially and temporally complex, and al-
M

0.0! most always affect the entire heart [31]. Until recently, whole
heart simulations were virtually impossible because of a lack
of sufficient image resolution and computational power. Re-
+2.0! cent advances in non-invasive imaging and computer simula-
D

tion now allows us to create fully three-dimensional models of


0! 100! 200! 300! 400! the entire human heart to explore the interplay between electri-
time [ms]!
TE

cal and mechanical fields under healthy, diseased, and treatment


Figure 9: Temporal evolution of long-axis shortening. During ventricular ejec- conditions [20]. With these goals in mind, we have designed a
tion, the distance between apex and base decreases rapidly as the ventricles basic prototype model for excitation-contraction coupling in the
shorten by approximately 7mm. During ventricular filling, the long axis gradu- human heart.
ally returns to its initial length as the heart muscle relaxes.
EP

In this proof-of-concept study, we have demonstrated that it


120! is feasible to model whole heart function within a single, uni-
ES! fied finite-element based modeling environment. First, we have
100! shown that our whole heart simulation predicts spatio-temporal
C

ventricular profiles of the electrical potential, the mechanical deformation,


80! ejection! and the muscle fiber strain, which agree nicely with clinical ob-
pressure [mmHg]!

servations and engineering intuition [48]. Throughout the car-


AC

isovolumetric isovolumetric diac cycle, the electrical potential varies between -80mV and
60! relaxation! contraction!
+20mV, the mechanical deformation takes values in the 10mm
40! ventricular! regime, and the maximum muscle fiber contraction is in the
filling! order of 20%, see Figure 7. Second, we have illustrated how
20! our heart muscle cells locally depolarize rapidly from the rest-
ing state at -80mV to the excited state at +20mV and depolar-
ED!
0! ize gradually from +20mV back to -80mV [15], see Figure 8.
80! 85! 90! 95! 100! 105! Last, we have globally extracted metrics of cardiac function,
volume [mL]! long-axis shortening and pressure-volume loops, see Figures
Figure 10: Pressure-volume loop of the human heart with characteristic phases 9 and 10. Our long-axis shortening of approximately 7.0mm
of ventricular filling, isovolumetric contraction, ventricular ejection, and iso- agrees nicely with previous simulations of a bi-ventricular hu-
volumetric relaxation. The enclosed area characterizes the work performed man heart model, which predicted a shortening of 7.6mm [48]
throughout the cardiac cycle.
and with clinical observations [39]. Our pressure-volume loop
8
ACCEPTED MANUSCRIPT

with left ventricular volumes between 83mL and 103mL and stretch-induced changes in conductivity, whereas a mechani-
pressures between 5mmHg and 118mmHg lies nicely within cally sensitive source f φ (F) would mimic the effect of stretch-
the clinically expected range [4, 8]. Our current run time for activated ion channels [33]. Neglecting these effects allows us
the electro-mechanical simulation of an entire cardiac cycle is to solve the electrical and mechanical problems in a decoupled
94 minutes on a standard 16 CPU machine. way. While we ultimately aim to solve the coupled problem of
electro-mechanics in equations (18) fully monolithically, here,
6.1. Limitations and future perspectives we solve the electrical module (18.1) using Abaqus/Standard
Our geometric representation of the human heart is promising 6.13 and the mechanical module (18.2) using Abaqus/Explicit
and our first simulations are encouraging, both from an engi- 6.13 [1].
neering and clinical perspective. Yet, our current model has a For the fluid module, we have assumed a simplified resistance

PT
few limitations, which can be addressed with modular changes model of Windkessel type [13]. While some approaches sug-
and refinements. In particular, our ongoing work aims to im- gest to improve this model with additional capacitors or time-
plement the following enhancements: varying resistances [43], our ultimate goal is to replace the com-
For the electrical module, as a first step, we have assumed that partment approach with a real fluid simulation. A fully coupled

RI
the cellular response is homogeneous across the heart. Recent fluid-structure interaction model of the human heart is highly
studies have shown that cellular heterogeneity is critical to ac- desirable, albeit hugely challenging, and the major research
curately represent the electrical activation sequence of the heart: thrust in numerous groups around the world. Including fluid

SC
The first regions to depolarize are the last to repolarize [27]. flow would allow us to predict shear stresses on the myocardial
This is true for both regional and transmural activation and can wall, and more importantly, on the four heart valves, throughout
be addressed via different cell types or via a spatially varying re- the entire cardiac cycle [21]. This presents tremendous oppor-
fractoriness [25]. In addition, cellular conduction varies hugely tunities to better understand the mechanisms of valvular disease

U
between standard cardiac muscle cells and Purkinje fiber cells, and optimize their treatment in the form of valve repair or re-
which are fast-conducting cells that transmit the electrical sig- placement, either through open heart surgery or minimally in-
AN
nal rapidly from the atrioventricular node to the apex [27]. We vasive intervention [14].
can incorporate this fast conduction, which is critical for bot-
tom up excitation, by adding one-dimensional cable elements 6.2. Concluding Remarks
across the inner wall [29]. We have presented a proof-of-concept simulator for cardiac ex-
citation and contraction in the human heart. Using human com-
M

For the mechanical module, probably the most significant limi-


tation is the parameter identification, which has been performed puter tomography and magnetic resonance images, we have cre-
on explanted tissue samples. Recent studies have indicated that ated a whole heart model with all four chambers, connected
in vivo properties can be up to four orders of magnitude dif- through four valves. The coordinated opening and closing of
D

ferent from ex vivo properties [38]. Ideally, a series of in vivo these valves regulates the filling of the chambers; the controlled
experiments should be designed around calibrating these pa- interplay of electrical and mechanical fields coordinates their
ejection. To simulate the blood flow from chamber to cham-
TE

rameters in the beating heart [46]. While the passive properties


are relatively easy to access, identifying the active properties ber, we have adopted a classical resistance-based Windkessel
can be challenging. Here we have assumed a phenomenolog- model. To simulate passive filling and active contraction, we
ical representation of active contraction through equations (9) have implemented a two-field finite element formulation based
on coupled electrical and mechanical fields. We have shown
EP

and (10). Integrating cellular phenomena such as calcium re-


lease, actin-myosin sliding, and cross bridging could provide that our model is capable of predicting the spatio-temporal
a more mechanistic representation of active force generation evolution of electrical potentials and mechanical deformation
[34]. The current trend is to replace active stress with active across the heart. From these, we have extracted two common
metrics of cardiac function, long-axis shortening and pressure-
C

strain [19, 37], a property that is easier to interpret, easier to


bound with physically meaningful values, easier to link to tis- volume loops, which agreed well with clinical observations.
sue microstructure [41], and easier to measure overall. Our ultimate goal is to employ our human heart simulator to
AC

For the interaction between the electrical and mechanical fields, probe landscapes of clinical parameters, and guide device de-
we have assumed that coupling is primarily uni-directional: sign and treatment planning in cardiac diseases of the aortic,
Changes in the electrical field φ induce major changes in the pulmonary, tricuspid, or mitral valve such as stenosis, regurgi-
mechanical field ϕ, while changes in the mechanical field only tation, or prolapse, and in other forms of cardiac dysfunction.
induce marginal or no changes in the electrical field. The en-
tries of the coupling matrices Kϕφ φϕ
JK and KIL and in equations Acknowledgements
(19.2) and (19.3) provide quantitative insight into the nature
This work was performed as part of the Living Heart Project. The
of these coupling effects, which manifest themselves in the ac- authors thank Dr. Julius M. Guccione, Dr. Manuel K. Rausch, Dr.
tive stress Pact , the electrical flux q, and the electrical source Jonathan Wong, and Dr. Martin Genet for stimulating discussions.
f phi . Our current model neglects the effects of mechanical de- Ellen Kuhl acknowledges support by the National Science Foundation
formation on the electrical field [16]. These phenomena, which CAREER award CMMI 0952021, by the National Science Foundation
are commonly believed to be of minor importance, come in INSPIRE grant 1233054, and by the National Institutes of Health grant
two flavors: A mechanically sensitive flux q(F) would mimic U54 GM072970.
9
ACCEPTED MANUSCRIPT

References [27] Keener, J., Sneyd, J., 2004. Mathematical Physiology. Springer Science
and Business Media.
[1] Abaqus 6.13. Analysis User’s Manual. 2013. Simulia. Dassault Systèmes. [28] Klepach, D., Lee, L.C., Wenk, J.F., Ratcliffe, M.B., Zohdi, T.I., Navia,
[2] American Heart Association. 2014. Heart disease and stroke statistics J.L., Kassab, G.S., Kuhl, E., Guccione, J.M., 2012. Growth and remodel-
2014 update. Circulation. 129:e28e292. ing of the left ventricle: A case study of myocardial infarction and surgi-
[3] Aliev, R.R., Panfilov, A.V., 1996. A simple two-variable model of cardiac cal ventricular restoration. Mech. Res. Comm. 42:134-141.
excitation. Chaos. 7,293-301. [29] Kotikanyadanam, M., Göktepe, S., Kuhl, E., 2010. Computational mod-
[4] Berberoglu, E., Solmaz, H.O., Göktepe, S., 2014. Computational model- eling of electrocardiograms - A finite element approach towards cardiac
ing of coupled cardiac electromechanics incorporating cardiac dysfunc- excitation. Int. J. Num. Meth. Biomed. Eng. 26:524-533.
tions. Euro. J. Mech. A/Solids. this issue. [30] Krishnamoorthi, S., Sarkar, M., Klug, W.S., 2013. Numerical quadrature
[5] Berne, R.M., Levy, M.N., 2001. Cardiovascular Physiology. The Mosby and operator splitting in finite element methods for cardiac electrophysil-
Monograph Series. ogy. Int. J. Num. Meth. Biomed. Eng. 29:1243-1266.

PT
[6] Bonet, J., Wood, R.D., 1997. Nonlinear Continuum Mechanics for Finite [31] Kumar, V., Abbas, A.K., Fausto, N., 2005. Robbins and Cotran Pathologic
Element Analysis. Cambridge University Press. Basis of Disease. Elsevier Saunders.
[7] Braunwald E. 1997. Heart Disease: A Textbook of Cardiovascular [32] Lee, L.C., Wenk, J.F., Zhong, L., Klepach, D., Zhang, Z.H., Ge, L., Rat-
Medicine. W.B. Saunders Company. cliffe, M.B., Zohdi, T.I., Hsu, E., Navia, J.L., Kassab, G.S., Guccione,
[8] Burghoff, D., 2013. Pressure-volume loops in clinical research. J. Am. J.M., 2013. Analysis of patient-specific surgical ventricular restoration:

RI
College Cardiology. 62:1173-1176. Importance of an ellipsoidal left ventricular geometry for diastolic and
[9] Dal, H., Göktepe, S., Kaliske, M., Kuhl, E., 2012. A fully implicit fi- systolic function. J. Appl. Phys. 115:136-144.
nite element method for bidomain models of cardiac electrophysiology. [33] Markhasin, V.S., Solovyova, O., Katsnelson, L.B., Protsenko, Y., Kohl,
Comp. Meth. Biomech. Biomed. Eng. 15,645-656. P., Noble D., 2003. Mechano-electric interactions in heterogeneous my-

SC
[10] Dal, H., Göktepe, S., Kaliske, M., Kuhl, E., 2013. A fully implicit finite ocardium: Development of fundamental experimental and theoretical
element method for bidomain models of cardiac electromechanics. Comp. models. Prog. Biophys. Mol. Biol. 82:207-220.
Meth. Appl. Mech. Eng. 253:323-336. [34] Murtada, S.C., Arner, A., Holzapfel, G.A., 2012. Experiments and
[11] Eriksson, T.S.E., Prassl, A.J., Plank, G., Holzapfel, G.A., 2013. Modeling mechanochemical modeling of smooth muscle contraction: Significance
the dispersion in electromechancially coupled myocardium. Int. J. Num. of filament overlap. J. Theor. Bio. 297:176-186.
Meth. Biomed. Eng. 29:1267-1284.

U
[35] Nagumo, J., Arimoto, S., Yoshizawa, S., 1962. An active pulse transmis-
[12] Fitzhugh, R., 1961. Impulses and physiological states in theoretical mod- sion line simulating nerve axon. Proc. Inst. Radio Eng. 50:2061-2070.
els of nerve induction. Biophys. J. 1:455-466. [36] Nash, M.P., Panfilov, A.V., 2004. Electromechanical model of excitable
AN
[13] Frank, O., 1899. Die Grundform des arteriellen Pulses. Zeitung für Biolo- tissue to study reentrant cardiac arrhythmias. Progr. Biophys. Mol. Bio.
gie. 37:483-586. 85:501-522.
[14] Gessat, M., Hopf, R., Pollok, T., Russ, C., Frauenfelder, T., Sundermann, [37] Pezzuto, S., Ambrosi, D., Quarteroni, A., 2014. An orthotropic active-
S.H., Hirsch, S., Mazza, E., Szekely, G., Falk, V., 2014. Image-based strain numerical model for the mechanics of the myocardium. Euro. J.
mechanical analysis of stent deformation: Concept and exemplary imple- Mech. A/Solids. this issue.
mentation for aortic valve stents IEEE Trans. Biomed. Eng. 61:4-15.
M

[38] Rausch, M.K., Kuhl, E., 2013. On the effect of prestrain and residual
[15] Göktepe, S., Kuhl, E., 2009. Computational modeling of electrophysiol- stress in thin biological membranes. J. Mech. Phys. Solids. 61:1955-1969.
ogy: A novel finite element approach. Int. J. Num. Meth. Eng. 79:156- [39] Robers, W.J., Shapiro, E.P., Weiss, J.L., Buchalter, M.B., Rademakers,
178. F.E., Weisfeldt, M.L., Zerhouni, E.A., 1991. Quantification of and cor-
[16] Göktepe, S., Kuhl, E., 2010. Electromechanics of the heart - A unified relation for left-ventricular systolic long-axis shortening by magnetic-
D

approach to the strongly coupled excitation-contraction problem. Comp. resonance tissue tagging and slice isolation. Circulation. 84: 721-731.
Mech. 45:227-243. [40] Rossi, S., Ruiz-Baier, R., Pavarino, L., Quarteroni, A., 2012. Orthotropic
[17] Göktepe, S., Wong, J., Kuhl, E., 2010. Atrial and ventricular fibrillation active strain models for the numerical simulation of cardiac biomechan-
TE

- Computational simulation of spiral waves in cardiac tissue. Arch. Appl. ics. Int. J. Num. Meth. Biomed. Eng. 28:761-788.
Mech. 80:569-580. [41] Rossi, S., Lassila, T., Ruiz-Baier, R., Sequeira, A., Quarteroni, A., 2014.
[18] Göktepe, S., Acharya, S.N.S., Wong, J., Kuhl, E., 2011. Computational Thermodynamically consistent orthotropic activation model capturing
modeling of passive myocardium. Int. J. Num. Meth. Biomed. Eng. 27:1- ventricular systolic wall thickening in cardiac electromechanics. Euro. J.
12. Mech. A/Solids. this issue.
EP

[19] Göktepe, S., Menzel, A., Kuhl, E., 2014. The generalized Hill model: [42] Shmukler, M., 2004. Density of Blood. The Physics Factbook.
A kinematic approach towards active muscle contraction. submitted for [43] Smith, B.W., 2004. Minimal haemodynamic modelling of the heart and
publication. circulation for clinical application. Dissertation. University of Canterbury,
[20] Gonzales, M.J., Sturgeon, G., Krishnamurthy, A., Hake, J., Jonas, R., Christchurch, New Zeland.
Stark, P., Rappel, W.J., Narayan, S.M., Zhang, Y., Segars, W.P., McCul- [44] Trayanova, N.A., 2011. Whole-heart modeling: Applications to cardiac
C

loch, A.D., 2013. A three-dimensional finite element model of human ar- electrophysiology and electromechanics. Circ. Res. 108:113-128.
trial anatomy: New methods for cubic Hermite meshes with extraordinary [45] Trayanova, N.A., Constantino, J., Gurev, V., 2011. Electromechanical
ventricles. Med. Im. Anal. 17:525-537. models of the ventricles. Am. J. Phys. Heart Circ. Phys. 301:H279-H286.
AC

[21] de Hart, J., Peters, G.W.M., Schreurs, P.J.G., Baaijens, F.P.T., 2003. A [46] Tsamis, A., Bothe, W., Kvitting, J.P., Swanson, J.C., Miller, D.C., Kuhl,
three-dimensional computational analysis of fluid-structure interaction in E., 2011. Active contraction of cardiac muscle: In vivo characterization of
the aortic valve. J Biomech. 36:103-112. mechanical activation sequences in the beating heart. J. Mech. Behavior
[22] Holzapfel, G.A., 2000. Nonlinear Solid Mechanics: A Continuum Ap- Biomed. Mat. 4:1167-1176.
proach for Engineering. John Wiley & Sons. [47] Walker, J.C., Ratcliffe, M.B., Zhang, P., Wallace, A.W., Fata, B., Hsu,
[23] Holzapfel, G.A., Ogden, R.W., 2009. Constitutive modelling of passive E.W., Saloner, D., Guccione, J.M., 2005. MRI-based finite-element anal-
myocardium: A structurally based framework for material characteriza- ysis of left ventricular aneurysm. Am. J. Physiol. Heart Circ. Physiol. 289:
tion. Phil. Trans. A. 367:3445-3475. H692-H700.
[24] Hunter, P.J., Pullan, A.J., Smaill, B.H., 2003. Modeling total heart func- [48] Wong, J., Göktepe, S., Kuhl, E., 2013. Computational modeling of
tion. Ann. Rev. Biomed. Eng. 5:147-177. chemo-electro-mechanical coupling: A novel implicit monolithic finite
[25] Hurtado, D.E., Kuhl, E. Computational modeling of electrocardiograms: element approach. Int. J. Num. Meth. Biomed. Eng. 29:1104-1133.
Repolarization and T-wave polarity in the human heart. Comp. Meth. [49] Wong, J., Kuhl, E., 2014. Generating fiber orientation maps in hu-
Biomech. Biomed. Eng. doi:10.1080/10255842.2012.729582. man heart models using Poisson interpolation. Comp. Meth. Biomech.
[26] Hurtado, D.E., Henao, D. Gradient flows and variational principles for Biomed. Eng. doi:10.1080/10255842.2012.739167.
cardiac electrophysiology: Toward efficient and robust numerical simula- [50] Zygote Media Group, Inc., 2013. The Zygote Solid 3D Heart Model.
tions of the electrical activity of the heart. Comp. Meth. Appl. Mech. Eng.
273:238-254.

10

Вам также может понравиться