Download as pdf or txt
Download as pdf or txt
You are on page 1of 24

Computer Methods in Biomechanics and Biomedical

Engineering

ISSN: 1025-5842 (Print) 1476-8259 (Online) Journal homepage: https://www.tandfonline.com/loi/gcmb20

A finite element model of the human left


ventricular systole

F. Dorri , P. F. Niederer & P. P. Lunkenheimer

To cite this article: F. Dorri , P. F. Niederer & P. P. Lunkenheimer (2006) A finite element model
of the human left ventricular systole, Computer Methods in Biomechanics and Biomedical
Engineering, 9:5, 319-341, DOI: 10.1080/10255840600960546

To link to this article: https://doi.org/10.1080/10255840600960546

Published online: 25 Jan 2007.

Submit your article to this journal

Article views: 2065

View related articles

Citing articles: 2 View citing articles

Full Terms & Conditions of access and use can be found at


https://www.tandfonline.com/action/journalInformation?journalCode=gcmb20
Computer Methods in Biomechanics and Biomedical Engineering,
Vol. 9, No. 5, October 2006, 319–341

A finite element model of the human left ventricular systole


F. DORRI†*, P. F. NIEDERER† and P. P. LUNKENHEIMER‡

†Institute for Biomedical Engineering, University and ETH Zurich, Gloriastrasse 35, CH-8092 Zurich, Switzerland
‡Experimental Thoracovascular and Cardiac Surgery, University of Munster, Domagkstrasse 11, D-48149 Munster, Germany

(Received 20 July 2005; in final form 8 August 2006)

Local wall stress is the pivotal determinant of the heart muscle’s systolic function. Under in vivo
conditions, however, such stresses cannot be measured systematically and quantitatively. In contrast,
imaging techniques based on magnetic resonance (MR) allow the determination of the deformation
pattern of the left ventricle (LV) in vivo with high accuracy. The question arises to what extent
deformation measurements are significant and might provide a possibility for future diagnostic
purposes.
The contractile forces cause deformation of LV myocardial tissue in terms of wall thickening,
longitudinal shortening, twisting rotation and radial constriction. The myocardium is thereby
understood to act as a densely interlaced mesh. Yet, whole cycle image sequences display a distribution
of wall strains as function of space and time heralding a significant amount of inhomogeneity even
under healthy conditions. We made similar observations previously by direct measurement of local
contractile activity. The major reasons for these inhomogeneities derive from regional deviations of the
ventricular walls from an ideal spheroidal shape along with marked disparities in focal fibre orientation.
In response to a lack of diagnostic tools able to measure wall stress in clinical routine, this
communication is aimed at an analysis and functional interpretation of the deformation pattern of an
exemplary human heart at end-systole. To this end, the finite element (FE) method was used to simulate
the three-dimensional deformations of the left ventricular myocardium due to contractile fibre forces at
end-systole. The anisotropy associated with the fibre structure of the myocardial tissue was included in
the form of a fibre orientation vector field which was reconstructed from the measured fibre trajectories
in a post mortem human heart. Contraction was modelled by an additive second Piola –Kirchhoff active
stress tensor.
As a first conclusion, it became evident that longitudinal fibre forces, cross-fibre forces and shear
along with systolic fibre rearrangement have to be taken into account for a useful modelling of systolic
deformation. Second, a realistic geometry and fibre architecture lead to typical and substantially
inhomogeneous deformation patterns as they are recorded in real hearts. We therefore, expect that the
measurement of systolic deformation might provide useful diagnostic information.

Keywords: Heart; Myocardium; Contraction; Systole

1. Introduction It is the task of mathematical modelling to bridge the


gap between ventricular morphology and mechanical
Recent developments in cardiac imaging, primarily by function. To this end, a number of theoretical approaches
way of magnetic resonance (MR), but also with ultrasound and numerical models have been put forward to assess
and X-ray allow for an advanced analysis and documen- cardiodynamics theoretically and to calculate the stress
tation of ventricular wall motion under in vivo conditions. and strain distribution in the ventricular wall (Chadwick
The spatial and temporal resolution of these imaging 1982, Arts et al. 1982, Nielsen et al. 1991, Hunter et al.
methods has in particular been improved to an extent that 1992, Huyghe et al. 1992, Bovendeerd et al. 1992, 1994,
the assessment of regional wall dynamics has become a 1996, Nash and Hunter 2000, Shoucri 2000, Guccione
realistic option in clinical routine. These developments et al. 2001a,b). These models have generally been based
provide a basis for a renewed and far more detailed on idealised, geometrical representations of animal and
interpretation of the driving structures that keep the human hearts; likewise, the anisotropic structure associ-
ventricular wall moving than has been possible hitherto. ated with the myocardial fibre architecture has been

*Corresponding author. Email: dorri@biomed.ee.ethz.ch

Computer Methods in Biomechanics and Biomedical Engineering


ISSN 1025-5842 print/ISSN 1476-8259 online q 2006 Taylor & Francis
http://www.tandf.co.uk/journals
DOI: 10.1080/10255840600960546
320 F. Dorri et al.

smoothed and modelled in an idealised fashion, measurements along with a mathematical heart model
disregarding local details. can be utilised for diagnostic purposes, in particular with
A scrutiny of the human myocardium reveals that it respect to the ventricular fibre pattern and wall stress
represents an anisotropic structure whereby the myocytes distribution.
are multi-branched and organised in bundles which are This work is devoted to an analysis of the deformation
interconnected by a complex hierarchy of connective pattern of an exemplary human heart under the influence
tissue (Feneis 1943, Hort 1960, Streeter and Basset 1966, of systolic contraction forces. To this end, the geometry of
Streeter et al. 1969, Spotnitz et al. 1974, LeGrice et al. the epi- and endo-cardial surfaces along with the fibre
1995, Hanley et al. 1999). In search of a functional direction field was derived from a post mortem human
understanding, various tertiary structural concepts heart. A continuum mechanics’ approach, i.e. a finite
regarding the ventricular architecture have been postu- element (FE) formulation was chosen for the calculation
lated in the past, thereby in particular, tractus (Mall of systolic contraction. As constitutive model, an
1911), bands, ropes (Torrent-Guasp et al. 1997, 2001), extension of the formulation of Lin and Yin (1998), for
helices (Ingels 1997) or lamellae (LeGrice et al. 1995, transversely isotropic materials was used, and the systolic
Young et al. 1998) although corresponding histological stimulation was modelled by the addition of an active
evidence is only partially conclusive. Two further, often component to the second Piola – Kirchhoff stress tensor.
tacitly made assumptions associated with such concepts As representative for the process the end-systolic state was
are that the wall structure is largely the same throughout considered and compared with measurements within the
the ventricle and that the myocardial fibres are aligned limits given by the fact that we analysed an individual case
almost completely in a tangential direction with respect associated with individual aberrations.
to the epi-cardial surface. Although the main fibre The chosen constitutive model did not allow to include
arrangement is found to exhibit a typical gross the effect of systolic fibre rearrangement, i.e. the effect of
architecture, which is similar in all mammalian hearts the cleavage clefts as well as the influence of the
examined so far, there are marked local and individual hierarchical organisation of the connective tissue or the
variations throughout the ventricular wall. Likewise, “accordion-like” contraction of the myocardial tissue
short axis sections of a healthy left ventricle (LV) are mentioned recently by Harrington et al. (2005), explicitly.
mostly round but with numerous local irregularities. Nevertheless, by adjusting the compressibility of the
There is furthermore evidence that countless, very short material, this shortcoming could partially be compensated.
cleavage clefts are interspersed throughout the myocar- Since the model was derived from one human heart, it
dium which facilitate some relative movement between is exemplary, represents a particular case and not a
fibre strands, in particular fibre rearrangement during population average. An extrapolation to other hearts is not
systolic contraction (LeGrice et al. 1997, Costa et al. a priori possible, however, we consider the chosen heart
1999, Usyk et al. 2000). The geometry, size and as representative with respect to the extent of the local
orientation of these cleavage planes are highly variable; variations which are to be expected in the systolic
they are found to exhibit a wide range of variability in contraction pattern due to the individual characteristics of
spatial alignment. a healthy human heart.
From a geometrical point of view, the main features of
systolic ventricular contraction are wall thickening,
longitudinal shortening, rotational wringing and radial 2. Material behaviour
constriction of the myocardium. These patterns are,
however, found to be subjected to appreciable local Myocardial tissue is anisotropic and weakly compressible
variations; in particular, wall thickening is quite because of intra- and extra-vascular fluid displacements
inhomogeneous throughout the ventricle (Chadwick et al. occurring during systole (Yin et al. 1996). The material
1989). Likewise, measured wall stress exhibits pro- behaviour of myocardium depends furthermore on time
nounced and characteristic variations (Lunkenheimer et al. and it changes during a heart cycle. Anisotropy, which is
2004). Mathematical models which have been presented due to the fibrous structure of the tissue, causes moreover
to date, however, have only partially been aimed at the directional dependency of the material behaviour on the
description and analysis of these features. applied loads. Transversely isotropic material models of
It is conceivable that stress distributions within the various functional forms and material parameters have
myocardium can be calculated from an inverse been widely used for the mathematical modelling of the
mathematical analysis (Moustakidis et al. 1999). A LV. The assumption of transverse isotropy for myocardial
prerequisite for such an analysis is a detailed knowledge tissue, i.e. the existence of a unique preferred direction
of the constitutive properties of myocardium including in the material, enables the deduction of a strain energy
the systolic contraction behaviour. MR techniques allow function directly from experimental data (Humphrey et al.
furthermore the determination of local fibre shortening 1990a,b, Novak et al. 1994). Recently, also orthotropic
during systole (Rademakers et al. 1994, MacGowan et al. material models for myocardium were proposed (Nash
1997). The question, therefore, arises to what extent and Hunter 2000, Usyk et al. 2000), but experimental data
a motion and deformation analysis based on MR are hardly sufficient to deduce the functional form of such
FE model of the human left ventricular systole 321

constitutive equations and to determine the associated The following values for the material parameters
material constants without further assumptions. ðg=cm2 Þ were chosen
Transversely isotropic composites can be modelled as
C 1 ¼ 27:89; C2 ¼ 66:20; C 3 ¼ 51:12;
being composed of fibre bundles which are embedded in a ð7Þ
homogeneous matrix (Spencer 1972, 1984) whereby the C 4 ¼ 40:12; C5 ¼ 0:0032
fibres are able to interact with the matrix. With C denoting
the right Cauchy – Green deformation tensor and N a unit These parameters correspond to the stiffest state under
vector which determines the anisotropy direction, a steady contraction according to the biaxial tests of Lin and
general form of the strain energy function in case of Yin (1998) and were used as baseline. A number of runs
incompressibility reads (Rivlin and Saunders 1951, were also made with other material parameters in order to
Humphrey et al. 1990) determine their influence.
Systolic intra- and extra-vascular fluid displacements
WðI 1 ; I 2 ; I 4 Þ ¼ W matrix ðI 1 ; I 2 Þ þ W fibres ðI 4 Þ give rise to a compressibility of the tissue. This is taken
into account by adding a volumetric part of the strain
þ W interaction ðI 1 ; I 2 ; I 4 Þ ð1Þ energy function
9K 1=3
I 1 ðCÞ ¼ tr C ð2Þ W Volumetric ¼ ðJ 2 1Þ2 ð8Þ
2
1  where J ¼ (det C)1/2. On the basis of the assumption that
I 2 ðCÞ ¼ ðtr CÞ2 2 tr C 2 ð3Þ the relative change in the volume of the ventricular wall
2
due to contraction is about 4% (Yin et al. 1996), a value
I3 ¼ 1 ð4Þ K ¼ 600 kPa ð9Þ
is appropriate. In order to take into account systolic fibre
I 4 ðC; NÞ ¼ NCN ð5Þ
arrangement, in turn, an effective compressibility
I1, . . . I4 are the invariant measures of deformations, K ¼ 70 kPa ð10Þ
whereby I3 ¼ 1 implies incompressibility.
was chosen. Accordingly, the total passive strain energy
In case of myocardial tissue, strain energy functions
function was written as
have mostly been based on two approaches (Weiss1994).
In the first approach, the symmetry properties of the W total ¼ W þ W Volumetric ð11Þ
structure are considered and the strain energy function is During the systolic phase, the active contraction forces
formulated in the local reference system reflecting the produced by the myocardial fibres increase gradually.
symmetries (Chuong and Fung 1986, Guccione et al. These forces were modelled in the form of additive
1991). Second, a strain energy function is introduced in components of the second Piola – Kirchhoff stress tensor
the global reference system which depends explicitly on which were added to the passive components in a stepwise
the fibre vector field N (identical with the anisotropy fashion described further down.
direction mentioned above). This field represents the
preferred direction of the material in each point of the
body (Humphrey et al. 1990a,b, Novak et al. 1994). 3. A representative fibre field N and the ventricular
We modelled the ventricular myocardium as a geometry
transversely isotropic material, composed of a weakly
compressible matrix and fibres (Spencer 1972, 1984), A healthy human heart in rigor mortis weighing about
whereby, the final constitutive behaviour consisted of a 500 g was chosen as a representative example. It was
passive as well as of an active component. perfused via the coronary arteries for 24 h with saline at a
The passive constitutive behaviour was formulated by pressure of 120 mm Hg. The perfusate was subsequently
choosing an appropriate form for Wtotal which was replaced by a 10% formaldehyde solution and the
consisted of a deviatoric (incompressible) part in the form perfusion was continued for another 24 h. The atria were
of equation (1) and a volumetric contribution describing trimmed down to the ventricular base and the heart was
the compressibility. For the deviatoric part, we used the submerged for two weeks in a 10% formaldehyde
functional form proposed by Lin and Yin (1998) for solution. The ventricles were then filled with Technovit
mammalian left ventricular myocardium under steady- (Haereus-Kulzer, Germany) such that mouldings of the
state barium stimulation for the simulation of systolic ventricular cavities were obtained. Together with the two-
excitation, viz., component resin a wooden rod was axially anchored in the
left ventricular cavity, which served to fix the heart in a jig
WðI 1 ; I 4 Þ ¼C 0 þ C4 ðI 1 2 3Þ þ C2 ðI 1 2 3Þ2 on top of an electromagnetic digitising tablet. After having
removed the epi-cardium together with the perivascular
þ C5 ðI 4 2 1Þ þ C3 ðI 4 2 1Þ2
fat, the geometries of the epi- and endo-cardial surfaces,
þ C1 ðI 1 2 3ÞðI 4 2 1Þ ð6Þ respectively, were determined. Subsequently, the SPOT
322 F. Dorri et al.

(fibre Strand Peel-Off Technique, see 7, Discussion) A geometrical model of the LV was produced from the
method was applied to prepare both ventricles including data sets containing about 2000 points representing the
the septum strand by strand, from base to apex and from endo-cardial and some 4500 points representing the epi-
epi-cardium to endo-cardium. During the entire process, a cardial surface (figure 2). For this purpose, Nonuniform
stepwise digitisation was performed by using a manual Rational B-Splines (NURBs) (Rogers 2001) were used to
stylus (figure 1). The data were processed to produce a connect the points in a smooth fashion (Raindrop
three-dimensional unit fibre vector field, N, which Geomagicw). A mesh of nine-node FE was created
described the fibre architecture for the entire LV. The whereby eight nodes were used to define hexahedral
procedure is described in detail by Lunkenheimer et al. elements while the ninth node was reserved for the
(1997), Cryer et al. (1997). pressure. More than 46,700 elements were necessary in

Figure 1. Fibre trajectories of the LV approximated as cubic splines from the digitised points outlining the myocardial fibre strands obtained from a post
mortem human heart in rigor mortis.
FE model of the human left ventricular systole 323

Figure 2. Geometry of the LV constructed from the data sets containing typically 2000 points representing the endo-cardial and some 4500 points
representing the epi-cardial surface.

order to include the desired level of anatomical details aorta, the atria and it is furthermore connected to the right
(MSC-Marc Mentatw). ventricle. Quantitative information on these boundary
The measured fibre trajectories obtained from SPOT conditions is not available, yet, in order to avoid undesired
were then introduced into the geometrical model such that rigid body displacements, appropriate boundary con-
the mesh covered the fibres entirely. For each element, the ditions have to be chosen in the FE formulation. In our
average fibre direction was determined as described by model, the motion of the basal elements was suppressed.
Dorri (2004). A further boundary condition derives from the
intracavital blood pressure acting on the endo-cardial
surface. The blood pressure in a real ventricle depends on
4. Boundary conditions time and location. In order to determine the spatial
distribution of the pressure as a function of time, the fluid
In the natural state, the myocardium is surrounded by the dynamics of the blood in the ventricle would have to be
pericardial sac. At the basis, the LV is restrained by the considered. However, since the pressure gradients in the
324 F. Dorri et al.

LV during ejection are mostly small (typically product of e01 and the vector 2k
, 10 mm Hg) in comparison with the absolute pressure,
we discarded local variations and assumed a uniform e02 ¼ 2e01 £ k ð16Þ
dependence of the pressure on time. The curve
As the long axis of the LV was chosen along the k-axis
of the global reference system ði; j; kÞ, e02 was directed
PLV ¼ 2944t 2 þ 245t 0 # t # 0:2 s ð12Þ
inward and perpendicular to the e01 and k axes. The third
direction, e03 , defining the coordinate zN followed from
represents a parabola which was considered to approxi-
mate a representative systolic pressure curve with a e03 ¼ e01 £ e02 ð17Þ
maximum of
The relation between the unit vectors e0m of the local
PMax ¼ 16:0 kPa tMax ¼ 0:13 s ð13Þ coordinate system (xF, yS, zN) and the unit vectors e n of the
global reference system (x, y, z) was given by
which corresponds to 120 mm Hg, the normal systolic transformation matrix Q
blood pressure for a healthy heart. 0 1
q11 q12 q13
This pressure was uniformly applied to the entire endo-
B C
cardial surface. Q¼B @ 21
q q22 q23 C
A ð18Þ
The potential impact of the right ventricular pressure on q31 q32 q33
septal deformation was modelled by the curve
with components
PRV ¼ 2237t 2 þ 62t 0 # t # 0:2 s ð14Þ
qmn ¼ e0m e n ð19Þ
with a maximum of
as follows
PMax ¼ 4:0 kPa tMax ¼ 0:13 s ð15Þ
{e0m } ¼ Q{e n } m; n ¼ 1; 2; 3 ð20Þ
which corresponds to 30 mm Hg. This pressure was Systolic contraction was modelled by defining the total
applied in the septal region. second Piola –Kirchhoff stress tensor S as the sum of the
In prescribing the intracavital pressures as boundary passive stress tensor S passive derived from the strain energy
condition, an indirect simulation is performed: In a real function (equation 11) and an additional active component
heart, the contractile forces create the intracavital S active :
pressure, while in our simulation the forces equilibrated
the pressure. This procedure is justified insofar as inertial S ¼ S passive þ S active ð21Þ
forces associated with wall motion can be neglected.
While for the FE calculation the tensor, S, had to be
known in the global coordinate system, the active
5. Modelling of active contraction component of the second Piola –Kirchhoff stress tensor
ðS0active Þ was first determined in the local coordinate
A global coordinate system (unit vectors i, j, k, associated system, because it depended on the local fibre direction, N.
coordinates x, y, z) was chosen such that the k-axis is To this end, the initial state corresponding to end-diastole
oriented along the long axis of the LV. In our geometry, was assumed as stress-free (eventual end-diastolic initial
which is characterised by pronounced local deviations stresses were assumed to be small in comparison with the
from a symmetric form this axis cannot be defined exactly, active stresses resulting at end-systole and were therefore
yet, this is of secondary importance only since the global ignored in our simulations). A fixed time step, Dt, was
coordinate system can be chosen arbitrarily. The j-axis is chosen and the intracavital pressure curve was subdivided
perpendicular to the k-axis and is oriented approximately accordingly (since the simulation was quasistatic, the time
in the anterior – inferior direction. The i-axis, finally, is step had no further influence). The components of the
perpendicular to both axes defined before (figure 2). active stress tensor were at each integration step subjected
A local coordinate system, xF, yS, zN for each element to an incremental increase ðS0 inc active Þ, which was chosen
was defined as follows. First, the fibre orientation was such that the intracavital pressure was overcome and could
determined in the middle point of each element as an be equilibrated at the end of each step, the criterion being
average vector of the measured neighbour fibre orien- quasistatic contraction (i.e. a sequence of incremental
tations (Dorri 2004), which we denoted by e01 . We equilibrium states). The assumption was thereby
identified this direction with positive values of xF, which implicitly made that all fibres exhibit the same contraction
therefore represented the axis along which the fibres were behaviour, i.e. all fibres were treated alike. In case of our
oriented in the middle point of the considered element. A thick-walled, anisotropic and irregular structure, this
second vector e02 , defining yS, was given as the vector integration scheme was however associated with a certain
FE model of the human left ventricular systole 325

amount of uncertainty which was basically inherent in the swelling of the geometry, moreover the twisting rotation
chosen procedure (prescribed intracavital pressure and was not simulated correctly.
wall stresses derived thereof instead of the physiologically The dense branching of the myocytes suggested that a
true opposite). Yet, parametric variation studies allowed shear component in the active stress tensor should be
us to ascertain that this uncertainty was within 10% of included; therefore, in a third step, an active stress tensor
the increment and as such substantially less than the S0active in the following form
inhomogeneity due to the fibre architecture. 0 1
In the development of a specific form of the active stress S0active ðF; FÞ 0 0
B C
tensor, a number of experimental facts were observed. S0active ¼ B 0 0 S0active ðS; NÞ C
First, from imaging of the heart (primarily by MRI and X- @ A
0 S0active ðN; SÞ S0active ðN; NÞ
ray ventriculography) can be concluded (Stuber et al.
1999), that there are four gross features regarding the ð23Þ
deformation process during the contraction of a healthy
LV. These features include wall thickening, longitudinal was used. Again, no quantitative information was
shortening, twisting rotation and radial constriction. available that would allow to choose the amount of the
Second, the amount of cross-fibre stresses has recently shear. We concluded from the simulations that the shear
been studied experimentally (Lin and Yin 1998). These component S0active ðS; NÞ should be between 3 and 7% of the
measurements along with other observations document fibre component S0active ðF; FÞ.
that fibre, cross-fibre and shear stresses develop We set the incremental values of the second Piola –
simultaneously in the myocardium as is to be expected Kirchhoff active stress tensor S0 inc
active as follows
due to the syncytial nature of the tissue (Mazhari and 0 1
McCulloch 1999, Usyk et al. 2000). 3 0 0
Accordingly, we conducted a consecutive analysis of inc B C
S0 active ¼ B
@ 0 0 0:1 A
C ð24Þ
the influence of the various components of the active stress
tensor in view of these effects. Thereby, there was no 0 0:1 1:8
influence of sarcomere length or the velocity of sarcomere
shortening as the simulation was essentially quasistatic. Here we assumed that the value of S0active ðN; NÞ is about
First calculations were performed with an active stress 60% of S0active ðF; FÞ, and the value of S0active ðS; NÞ is about
tensor which contained only one component S0active ðF; FÞ, 3% of S0active ðF; FÞ. All values of the stress components in
i.e. the component in the fibre direction was incrementally (equation 24) are given in kPa.
increased ðS0 inc In the global coordinate system, S active could finally be
active Þ along with the prescribed intracavital
pressure as described previously. As expected, we found determined from
that the component S0active ðF; FÞ alone could not produce a
realistic deformation pattern. In fact, a simulation of the S active ¼ Q T S0active Q ð25Þ
contraction with S0active ðF; FÞ only caused longitudinal
whereby the transformation matrix Q had to be
elongation of the LV instead of shortening, as was already
determined for each element.
reported in previous FE models (Huyghe et al. 1992,
Bovendeerd et al. 1994).
To simulate the longitudinal shortening, in a second
6. Results
step, a normal component S0active ðN; NÞ, i.e. an active stress
in cross-fibre direction zN was added. The relation between
6.1 Evaluation procedure
S0active ðF; FÞ and S0active ðN; NÞ is not straightforward,
however. Nevertheless, the measurements performed by In our geometrical model of the human LV which exhibits
Lin and Yin (1998) on barium-contracted rabbit no symmetry, all calculations were made in a global
myocardium had shown that S0active ðN; NÞ could lie Cartesian coordinate system. To assess theoretically,
between 20 and 62% of the S0active ðF; FÞ. Accordingly, estimated values of strains with available experimental
the cross-fibre stress was assumed to be proportional to the data we, therefore, calculated the components of the strain
longitudinal fibre stress. Upon application of an active tensor in the fibre-specific coordinate system xF, yS, zN.
stress tensor with fibre and cross-fibre components in the Thereby, it has to be noted that the directions of the xF, yS,
form zN axes change during the deformation process and had to
0 1 be updated at the end of each increment. As the
S0active ðF; FÞ 0 0 components of the strain tensor depend on the local
B C orientation of the coordinate system, care has to be
S0active ¼ B
@
0 0 0 C
A ð22Þ
exercised when estimated values of strain components are
0 0 S0active ðN; NÞ compared with reported experimental data from high
resolution MR imaging (Moore et al. 2000, Sinusas et al.
the longitudinal shortening and wall thickening could be 2001). The values given in the following represent
simulated but it was associated with some unphysiological extrapolations to coordinate systems which are oriented
326 F. Dorri et al.

similarly as those usually applied when symmetric Twisting rotation is counterclockwise when viewed from
approximations are used. the apex but irregularly distributed from the base to the
In contrast to principal and normal strains, shear strains apex on the endo-cardial and epi-cardial surfaces.
cannot be measured with sufficient precision to allow for a Experimental data (Moore et al. 2000), have shown that
useful comparison. Accordingly, we restricted our analysis this angle increases between basal and apical levels from
to principal and normal strains. To make a regional the epi-cardium (about 10 degrees) to the endo-cardium
comparison of the model with the results of tagged MRI (about 14 degrees). In this simulation twisting rotation was
measurements in the normal human LV, we estimated about 10 degrees on the anterior wall.
average values of strain components in the basal, The average radial displacement of the myocardial
equatorial and apical areas of the LV, respectively. midwall material points is directed inward throughout the
LV but with significant spatial variations. According to
recent experimental reports (see table 4 of Moore et al.
6.2 Model with effective compressibility
(2000)), it is smaller in the septum (about 2 3 ^ 2 mm)
In the model with an increased compressibility ðK ¼ than in the lateral and inferior walls (about 2 5 ^ 2 mm).
70 kPaÞ to take into account systolic fibre rearrangement It was found to be largest at the apical inferior wall
using (equation 24), the volume of the LV decreased (2 6.3 ^ 1.3 mm) and lowest at the apical anterior wall
during systolic contraction from the initial value of (2 2.4 ^ 1.1 mm). The radial displacement of the midwall
118 ml to the final value of 47 ml, which corresponds to material points in our model from base to apex and around
a stroke volume of 71 ml, i.e. the ejection fraction was the LV showed also an irregular distribution; it was about
about 0.6. An overall assessment of our simulations 2 2.5 ^ 1 mm and in some locations near the base
showed that the final value of the (F, F) component of practically negligible. This could partially be attributed to
the active stress tensor (after completion of 45 the boundary condition applied at the base.
increments in this simulation) was largely in agreement The results of the estimated average values of the
with theoretical estimations (maximum value of about principal strains in basal, equatorial and apical areas of the
135 kPa) made by other investigators (McCulloch et al. LV are summarised in table 1. In most regions the principal
1992, Guccione and McCulloch 1993, Hunter et al. strain EMax was found to vary around 0.20, EInt around
1998). 2 0.20 and EMin around 2 0.27 within a range of about
The von Mises stresses were found to vary in most 0.07. Figure 3 shows colour-coded maps of calculated
regions around 100 kPa within a range of about 50 kPa. principal and normal strains in a septal –lateral long-axis
Some extreme values of stress were recorded which section at end-systole, components of the strain tensor are
however can be attributed to singular irregularities referenced to the end-diastole. Estimated average values
associated with surface elements and which are not of of strain along fibre ðEff Þ and cross-fibre directions (Ess ,
significance because they are mostly due to numerical Enn ) are summarised in table 2. Figure 3 shows in addition
effects associated with the modelling. that in most regions Eff varied around 2 0.20, while Ess
Due to the compressibility, wall thickening was varied around 0.20 and Enn around 2 0.27 within a range
somewhat less (around 20% in most areas) in comparison of about 0.07. In this simulation calculated principal
with reported values in vivo. Longitudinal shortening, in strains, fibre and cross-fibre strains are generally smaller
turn, measured from epi- and endo-cardial surfaces to the than experimental measurements (see table 3 of Moore
base were equal and about 20%, this value is comparable et al. (2000), Tseng et al. (2000)). Figures 4 –6, show
with experimental measurements. In a healthy adult heart transmural variation of radial (ERR), circumferential (ECC)
longitudinal shortening from endo-cardial surface is more and longitudinal (ELL) strains in basal, equatorial and
than epi-cardial surface to the base. For the LV it is about apical areas of the LV from epi-cardium to endo-cardium.
15% measured from the epi-cardial surface (Stuber 1997). Estimated average values of radial, circumferential and

Table 1. Estimated average of principal strains (EMax, EInt, EMin) ðK ¼ 70 kPaÞ.

Strain Septal Anterior Lateral Inferior


EMax
Basal 0.27 ^ 0.07 0.20 ^ 0.07 0.27 ^ 0.07 0.13 ^ 0.07
Equatorial 0.33 ^ 0.07 0.20 ^ 0.07 0.27 ^ 0.07 0.20 ^ 0.07
Apical 0.33 ^ 0.07 0.07 ^ 0.07 0.13 ^ 0.07 0.20 ^ 0.07
EInt
Basal 20.13 ^ 0.07 20.20 ^ 0.07 20.20 ^ 0.07 20.07 ^ 0.07
Equatorial 20.20 ^ 0.07 20.20 ^ 0.07 20.20 ^ 0.07 20.20 ^ 0.07
Apical 20.20 ^ 0.07 20.13 ^ 0.07 20.27 ^ 0.07 20.20 ^ 0.07
EMin
Basal 20.33 ^ 0.07 20.27 ^ 0.07 20.27 ^ 0.07 20.27 ^ 0.07
Equatorial 20.27 ^ 0.07 20.33 ^ 0.07 20.27 ^ 0.07 20.33 ^ 0.07
Apical 20.27 ^ 0.07 20.33 ^ 0.07 20.27 ^ 0.07 20.27 ^ 0.07
FE model of the human left ventricular systole 327

Figure 3. Colour-coded maps of strain pattern in a septal–lateral long-axis section at end-systole ðK ¼ 70 kPaÞ. These figures show distribution of
strain along: (A) fibre direction F ðEff Þ; (B) cross-fibre direction S ðEss Þ; (C) cross-fibre direction N ðEnn Þ; (D –F) show principal strains ðEMax ; EInt ; EMin Þ.
Components of the strain tensor are referenced to the end-diastole.
328 F. Dorri et al.

Table 2. Estimated average of normal strains along fibre (Eff) and cross-fibre(Ess, Enn) directions ðK ¼ 70 kPaÞ.

Strain Septal Anterior Lateral Inferior


Eff
Basal 20.13 ^ 0.07 20.20 ^ 0.07 20.20 ^ 0.07 20.13 ^ 0.07
Equatorial 20.20 ^ 0.07 20.20 ^ 0.07 20.20 ^ 0.07 20.20 ^ 0.07
Apical 20.13 ^ 0.07 20.20 ^ 0.07 20.27 ^ 0.07 20.27 ^ 0.07
Ess
Basal 0.20 ^ 0.07 0.13 ^ 0.07 0.20 ^ 0.07 0.13 ^ 0.07
Equatorial 0.20 ^ 0.07 0.13 ^ 0.07 0.20 ^ 0.07 0.13 ^ 0.07
Apical 0.20 ^ 0.07 0.07 ^ 0.07 0.13 ^ 0.07 0.13 ^ 0.07
Enn
Basal 20.27 ^ 0.07 20.27 ^ 0.07 20.27 ^ 0.07 20.20 ^ 0.07
Equatorial 20.27 ^ 0.07 20.20 ^ 0.07 20.20 ^ 0.07 20.20 ^ 0.07
Apical 20.27 ^ 0.07 20.27 ^ 0.07 20.27 ^ 0.07 20.20 ^ 0.07

longitudinal strains are summarised in table 3. In most smaller than experimental measurements (see table 2 of
regions the ERR was found to vary around 0.20, ECC Moore et al. (2000), Tseng et al. (2000)).
around 2 0.20 and ELL around 2 0.27 within a range of
about 0.07. As such, these results were smaller in
6.3 Model with weakly compressible material
comparison to the model with a weekly compressible
material discussed in the next paragraph. Here, ECC is in In the model with weekly compressible material
agreement with measured values but ERR and ELL are ðK ¼ 600 kPaÞ, the results of the simulation using the

Figure 4. Variation of radial ðERR Þ, circumferential ðECC Þ and longitudinal ðELL Þ components of elastic strain (^0.07) in septal, anterior, lateral and
inferior of basal area from epi-cardium to endo-cardium ðK ¼ 70 kPaÞ.
FE model of the human left ventricular systole 329

Figure 5. Variation of radial ðERR Þ, circumferential ðECC Þ and longitudinal ðELL Þ components of elastic strain (^0.07) in septal, anterior, lateral and
inferior of equatorial area from epi-cardium to endo-cardium ðK ¼ 70 kPaÞ.

tensor according to (equation 24) are shown in (figure 7). degrees which is clearly represented on anterior wall
In this particular case, the volume of the LV decreased (figure 7).
from the initial value of 118 ml to the final value of 79 ml, The radial displacement of the midwall material points
which corresponds to a stroke volume of 39 ml i.e. ejection in this model from base to apex and around the LV showed
fraction was about 0.33. This low value for the ejection also an irregular distribution; but it was generally (too)
fraction will be commented later on in paragraph 7.2. All small (about 2 1.5 ^ 1 mm) and in some locations near
values of the von Mises stress in the pictures (figure 7) are the base practically negligible. Particularly the radial
given in kPa. These values are seen to vary in most regions inward motion of the apical wall was very small in
around 100 kPa within a range of about 50 kPa. Extreme comparison with experimental measurements.
values of stress for a number of elements were again The results of estimated average values of principal
confined to certain regions close to the surface. strains, summarised in table 4, are in good agreement with
Areas exhibiting high stresses are particularly well recent tagged MRI measurements of principal strains (see
apparent on the endo-cardial surface. As expected, wall table 3 of Moore et al. (2000), Tseng et al. (2000)). It can
thickening was inhomogeneously distributed; it is be seen that in most regions the principal strain EMax was
remarkable that in some locations maximal wall found to vary around 0.40, EInt around 2 0.13 and EMin
thickening was more than 40% as is measured in vivo. around 2 0.20 within a range of about 0.07. Figure 8
Longitudinal shortening, in turn, measured from epi- shows colour-coded maps of calculated principal and
cardial and endo-cardial surfaces to the base were equal normal strains in a septal –lateral long-axis section at end-
and about 5%, which is too small in comparison with systole. As expected, the strain pattern represented a
experimental measurements. In this model, the maximal heterogeneous distribution but with a limited range of
amount of the epi-cardial torsion was again about 10 fluctuations. Estimated average values of strain along fibre
330 F. Dorri et al.

Figure 6. Variation of radial ðERR Þ, circumferential ðECC Þ and longitudinal ðELL Þ components of elastic strain (^0.07) in septal, anterior, lateral and
inferior of apical area from epi-cardium to endo-cardium ðK ¼ 70 kPaÞ.

ðEff Þ and cross-fibre directions (Ess ,Enn ) are summarised showed that fibre shortening was more uniform than
in table 5. Figure 8 shows that in most regions Eff which cross-fibre or principal strains with a mean value of about
characterizes fibre shortening varied around 2 0.13, while 2 0.12 ^ 0.01 (Tseng et al. 2000).
Ess varied around 0.40 and Enn around 2 0.13 within a Figures 9 – 11 exhibit furthermore colour-coded maps of
range of about 0.07. These results are in agreement with strain distributions in basal, equatorial and apical short-
MRI measurements of myocardial fibre shortening in axis sections at end-systole. The essentially heterogeneous
human using diffusion-sensitive MR imaging which character of strain values with a limited band of fluctuations

Table 3. Estimated average of radial (ERR), circumferential (ECC) and longitudinal (ELL) strains ðK ¼ 70 kPaÞ.

Strain Septal Anterior Lateral Inferior


ERR
Basal 0.20 ^ 0.07 0.27 ^ 0.07 0.20 ^ 0.07 0.13 ^ 0.07
Equatorial 0.27 ^ 0.07 0.13 ^ 0.07 0.20 ^ 0.07 0.20 ^ 0.07
Apical 0.27 ^ 0.07 0.07 ^ 0.07 0.13 ^ 0.07 0.20 ^ 0.07
ECC
Basal 20.13 ^ 0.07 20.20 ^ 0.07 20.20 ^ 0.07 20.20 ^ 0.07
Equatorial 20.20 ^ 0.07 20.20 ^ 0.07 20.20 ^ 0.07 20.27 ^ 0.07
Apical 20.20 ^ 0.07 20.20 ^ 0.07 20.27 ^ 0.07 20.27 ^ 0.07
ELL
Basal 20.27 ^ 0.07 20.27 ^ 0.07 20.27 ^ 0.07 20.20 ^ 0.07
Equatorial 20.27 ^ 0.07 20.27 ^ 0.07 20.27 ^ 0.07 20.20 ^ 0.07
Apical 20.27 ^ 0.07 20.27 ^ 0.07 20.27 ^ 0.07 20.27 ^ 0.07
FE model of the human left ventricular systole 331

Figure 7. Simulation of the systolic contraction ðK ¼ 600 kPaÞ using the second Piola–Kirchhoff active stress tensor of equation (23) including
longitudinal fibre forces, cross fibre forces and shear.

could be determined also in short-axis sections. We of normal strains in septal, anterior, lateral and inferior
calculated also radial (ERR), circumferential (ECC) and from epi-cardium to endo-cardium. Additionally, esti-
longitudinal (ELL) strains in basal, equatorial and apical mated average values of radial, circumferential and
areas of the LV. Figures 12 –14 shows transmural variation longitudinal strains are summarised in table 6. In most

Table 4. Estimated average of principal strains (EMax, EInt, EMin) ðK ¼ 600 kPaÞ.

Strain Septal Anterior Lateral Inferior


EMax
Basal 0.40 ^ 0.07 0.40 ^ 0.07 0.40 ^ 0.07 0.27 ^ 0.07
Equatorial 0.60 ^ 0.07 0.40 ^ 0.07 0.47 ^ 0.07 0.33 ^ 0.07
Apical 0.60 ^ 0.07 0.20 ^ 0.07 0.40 ^ 0.07 0.47 ^ 0.07
EInt
Basal 20.07 ^ 0.07 20.13 ^ 0.07 20.13 ^ 0.07 20.07 ^ 0.07
Equatorial 20.13 ^ 0.07 20.13 ^ 0.07 20.13 ^ 0.07 20.13 ^ 0.07
Apical 20.13 ^ 0.07 20.07 ^ 0.07 20.13 ^ 0.07 20.13 ^ 0.07
EMin
Basal 20.27 ^ 0.07 20.20 ^ 0.07 20.20 ^ 0.07 20.20 ^ 0.07
Equatorial 20.20 ^ 0.07 20.20 ^ 0.07 20.27 ^ 0.07 20.20 ^ 0.07
Apical 20.27 ^ 0.07 20.20 ^ 0.07 20.20 ^ 0.07 20.20 ^ 0.07
332 F. Dorri et al.

Figure 8. Colour-coded maps of strain pattern in a septal–lateral long-axis section at end-systole ðK ¼ 600 kPaÞ. These figures show distribution of
strain along: (A) fibre direction F ðEff Þ; (B) cross-fibre direction S ðEss Þ; (C) cross-fibre direction N ðEnn Þ; (D –F) show principal strains ðEMax ; EInt ; EMin Þ.
Components of the strain tensor are referenced to the end-diastole.
FE model of the human left ventricular systole 333

Table 5. Estimated average of normal strains along fibre (Eff) and cross-fibre(Ess, Enn) directions ðK ¼ 600 kPaÞ.

Strain Septal Anterior Lateral Inferior


Eff
Basal 20.13 ^ 0.07 20.13 ^ 0.07 20.13 ^ 0.07 20.07 ^ 0.07
Equatorial 20.13 ^ 0.07 20.13 ^ 0.07 20.13 ^ 0.07 20.20 ^ 0.07
Apical 20.13 ^ 0.07 20.07 ^ 0.07 20.13 ^ 0.07 20.13 ^ 0.07
Ess
Basal 0.40 ^ 0.07 0.33 ^ 0.07 0.33 ^ 0.07 0.27 ^ 0.07
Equatorial 0.53 ^ 0.07 0.40 ^ 0.07 0.40 ^ 0.07 0.33 ^ 0.07
Apical 0.53 ^ 0.07 0.20 ^ 0.07 0.40 ^ 0.07 0.47 ^ 0.07
Enn
Basal 20.13 ^ 0.07 20.13 ^ 0.07 20.13 ^ 0.07 20.13 ^ 0.07
Equatorial 20.13 ^ 0.07 20.13 ^ 0.07 20.13 ^ 0.07 20.07 ^ 0.07
Apical 20.13 ^ 0.07 20.13 ^ 0.07 20.13 ^ 0.07 20.13 ^ 0.07

regions the ERR was found to vary around 0.33, ECC around hand-guided procedure and second, the peeling procedure
2 0.07 and ELL around 2 0.13 within a range of about 0.07. itself which isolates but randomly selected fibre
These results show that ERR and ELL are in good agreement populations of variable thickness, length and alignment,
with MRI measurements (see table 2 of Moore et al. which sequesters is currently called fibres or bundles. The
(2000), Tseng et al. (2000)), ECC on the other hand is larger totality of myocytes contained in such a bundle is
than measured values (around 2 0.20). excluded from their proper spatial analysis, and, more
prejudicial, the totality of the spatial contractile linkages
of the bundle with its surroundings abstains from being
displayed, except that one along which the strand
7. Discussion arbitrarily had been severed.
In consideration of all aspects, the SPOT technique,
7.1 Determination of the fibre field by the SPOT although associated with a number of drawbacks, can be
technique considered to expose the fibrous architecture realistically
Myocardial fibres are multiply branched. In histologic in that it can be assumed that a careful peeling, allowing
preparation, no beginnings and endings of fibre strands strands to follow the direction of least mechanical
can furthermore be found. Nevertheless, the tissue is resistance will expose the axis of mechanical anisotropy.
highly anisotropic and a preferred direction of the average
fibre orientation can clearly be discerned, because 7.2 Ejection fraction
myocytes form endless chains, which are densely meshed
by bridges to their paralleling neighbours. Due to this In a healthy adult heart, the ejection fraction, i.e. the ratio
syncytial nature of the tissue, however, the mean fibre of the stroke volume to the end diastolic volume under
orientation, which is considered to coincide with the axis resting condition, is about 0.60. Such a value was obtained
of anisotropy cannot be determined in a straightforward when an increased effective compressibility for the
fashion. Accordingly, whatever method one applies in the myocardial tissue was applied. In case of weekly
attempt to analyse the main alignment of myocardial compressible material, which is more realistic, however,
fibres, one has to destroy essential tissue structures. Mall’s the ejection fraction was considerably lower (33%). This
(1911), approach was to peel out of the ventricular muscle finding needs to be interpreted. One major mechanism for
distinct strands which he understood as being prominent the ejection of blood is wall thickening, which contributes
and functionally privileged parts of a continuum. By so around 45% to the LV stroke volume. Longitudinal
doing he accepted to destroy an unnumbered amount of shortening and radial inward motion of the wall
linkages which join the sequester with its bed. (circumferential shortening), in turn, have been estimated
In a previous study (Lunkenheimer et al. 1997), we to contribute around 55% of the stroke volume (Dumesnil
combined the peeling technique with an electromagnetic and Shoucri 1991). While wall thickening reached
digitising procedure by which the boundaries of the realistic values in our weekly compressible model, in
myocardial fibre strands, once isolated were compiled in particular longitudinal shortening was insufficient.
their spatial location and orientation. Then, the digitised Somewhat higher stroke volumes were reached by
strands were taken away to give access to deeper layers, a increasing the fibre forces. A further simulation was made
step which offered a rough access to the complex system with the following incremental second Piola– Kirchhoff
of interlayer connections. Finally, all digitised fibre active stress tensor
strands were compiled to reconstruct the original ventricle 0 1
9 0 0
in a computer simulation as composed of the totality of the B C
inc
individual strands in their spatial orientation. The S0 active ¼ B
@0 0 0:3 C
A: ð26Þ
drawback of our method was, first, the manual digitising 0 0:3 5:4
procedure which is susceptible to artefacts as any other
334 F. Dorri et al.

Figure 9. Colour-coded maps of strain pattern in a basal short-axis section at end-systole ðK ¼ 600 kPaÞ. These figures show distribution of strain
along: (A) fibre direction F ðEff Þ; (B) cross-fibre direction S ðEss Þ; (C) cross-fibre direction N ðEnn Þ; (D– F) show principal strains ðEMax ; EInt ; EMin Þ.
Components of the strain tensor are referenced to the end-diastole.

The number of increments and the bulk modulus K was corresponds to a stroke volume of 60 ml. As a result,
at the same time increased to 90 and 6 MPa, respectively, however, the stresses increased to unrealistically high
such that the relative change in the volume of the values while longitudinal shortening was still low.
ventricular wall due to contraction remained about 4%. In To assess furthermore, the effect of material parameters
this simulation the volume of the LV decreased from the in our calculations we used material constants of six other
initial value of 118 ml to the final value of 58 ml, which specimens suggested by Lin and Yin, (see table 1 of Lin
FE model of the human left ventricular systole 335

Figure 10. Colour-coded maps of strain pattern in an equatorial short-axis section at end-systole ðK ¼ 600 kPaÞ. These figures show distribution of
strain along: (A) fibre direction F ðEff Þ; (B) cross-fibre direction S ðEss Þ; (C) cross-fibre direction N ðEnn Þ; (D –F) show principal strains ðEMax ; EInt ; EMin Þ.
Components of the strain tensor are referenced to the end-diastole.
336 F. Dorri et al.

Figure 11. Colour-coded maps of strain pattern in an apical short-axis section at end-systole ðK ¼ 600 kPaÞ. These figures show distribution of strain
along: (A) fibre direction F ðEff Þ; (B) cross-fibre direction S ðEss Þ; (C) cross-fibre direction N ðEnn Þ; (D– F) show principal strains ðEMax ; EInt ; EMin Þ.
Components of the strain tensor are referenced to the end-diastole.
FE model of the human left ventricular systole 337

Figure 12. Variation of radial ðERR Þ, circumferential ðECC Þ and longitudinal ðELL Þ components of elastic strain (^0.07) in septal, anterior, lateral and
inferior of basal area from epi-cardium to endo-cardium ðK ¼ 600 kPaÞ.

and Yin (1998)). Resulting changes in ejection fraction reflected in our model. As a result, the mesh in the apical
and stress-strain pattern were minor. Using material region consisted of brick elements, which, under large
parameters (g/cm2) of (equation 27) increased the ejection curvature conditions, resist deformation generally. It
fraction to about 0.38, simultaneously higher strain should, however, be noted that most mathematical models
domains were reached in the wall without changing the have treated the apex as thick walled and only in a recent
longitudinal shortening. ventricular model in diastole it was attempted to construct
a more realistic structure of the apex (Stevens et al. 2003).
C 1 ¼ 213:14; C2 ¼ 17:05; C 3 ¼ 28:94; Due to the weak compressibility of the myocardium
ð27Þ extensive rearrangement of muscle fibres during systole is
C 4 ¼ 9:48; C5 ¼ 0:35 mandatory to yield sufficient systolic wall thickening,
circumferential and overall longitudinal shortening (Wald-
Material parameters are likely to influence the results of man et al. 1985, Rademakers et al. 1994, MacGowan et al.
simulation slightly but they do not change the pattern of 1997). Whereas the amount of fibre shortening is
deformation significantly. comparable from the base to the apex throughout the
Insufficient longitudinal shortening and radial inward myocardium, local wall thickening is quite inhomo-
displacement (circumferential shortening) were obviously geneous and increases significantly from sub-epi-cardium
the major causes for the low stroke volume in our weekly to sub-endo-cardium (Sabbah et al. 1981, Waldman et al.
compressible FE model. A first reason for this derives 1985, Rademakers et al. 1994). Realignment and
from the geometrical structure of apical wall, which interposition of fibres, causing an increase of the number
acuminates to a thin membrane-like structure. Due to of fibre cross-sections between epi- and endo-cardium
technical restrictions for the digitisation of the epi- during systole, appears to be necessary to produce these
cardium this geometrical feature was insufficiently effects, in particular longitudinal shortening (Hort 1960,
338 F. Dorri et al.

Figure 13. Variation of radial ðERR Þ, circumferential ðECC Þ and longitudinal ðELL Þ components of elastic strain (^0.07) in septal, anterior, lateral and
inferior of equatorial area from epi-cardium to endo-cardium ðK ¼ 600 kPaÞ.

Spotnitz et al. 1974). The intrinsic structure of the Measurements (Sabbah et al. 1981, Waldman et al. 1985,
myocardium including, cleavage clefts are major features Rademakers et al. 1994) showed furthermore that strain
to this end. It is therefore necessary to include some components, especially radial strain, in the sub-endo-
mechanism describing this properties; adapting material cardial region increases considerably which is a direct
constants or orthotropic material models alone cannot result of gliding or fibre rearrangement. The decreasing
change the pattern of deformation essentially (Usyk et al. strain components in our model in sub-endo-cardial areas,
2000). shown in figures 4 –6, 12– 14, can be attributed to the fact
The simulation using a constitutive law with a relatively that local curvature increases from epi-cardium to endo-
high effective compressibility showed that within the cardium and, as a result, the elements of the mesh in sub-
framework of our model, systolic interposition of fibres endo-cardial region resist deformation more than the
can artificially be modelled by increasing compressibility elements in sub-epi-cardial region.
of the material. Comparison of tables 1– 3 with tables 4–
6, as well as figures 4 – 6 with figures 12 –14, respectively,
show that assumption of compressibility of the myocar- 7.3 Comparison with other models
dium increases ejection fraction. This increase was During the last decades a number of FE models have been
provided essentially by increasing circumferential and put forward to assess cardiodynamics theoretically and to
longitudinal shortening, but it was accompanied by calculate the stress and strain distribution in the
reduction of other strain components such that important ventricular wall. These models have mostly analysed the
measures like wall thickening and fibre shortening were end-diastolic state and were based on idealised and
on their limits of the measured experimental range. This smoothed geometrical representations, furthermore, ani-
emphasises that the assumption of compressibility cannot mal hearts (dogs, pigs, and rats) were often modelled
replace the effect of fibre rearrangement completely. because available measurements are more abundant than
FE model of the human left ventricular systole 339

Figure 14. Variation of radial ðERR Þ, circumferential ðECC Þ and longitudinal ðELL Þ components of elastic strain (^0.07) in septal, anterior, lateral and
inferior of apical area from epi-cardium to endo-cardium ðK ¼ 600 kPaÞ.

in humans. A typical feature common to many models of system could then be defined with the help of the
systolic contraction (Horowitz et al. 1986, Huyghe et al. symmetry properties in the radial, circumferential and
1992, McCulloch et al. 1992, Bovendeerd et al. 1992, longitudinal directions. As a further simplification, the
1994, Guccione and McCulloch 1993, Guccione et al. fibre-specific coordinate system xF, yS, zN was defined
1995, Nash and Hunter 2000, Usyk et al. 2000) is the such that fibres were assumed as parallel to the epi-cardial
application of a symmetric geometry such as a thick surface. Another assumption was that active stresses
ellipsoid of revolution. A local cardiac-specific coordinate develop only along tangential fibre directions. These

Table 6. Estimated average of radial (ERR), circumferential (ECC) and longitudinal (ELL) strains ðK ¼ 600 kPaÞ.

Strain Septal Anterior Lateral Inferior


ERR
Basal 0.40 ^ 0.07 0.33 ^ 0.07 0.33 ^ 0.07 0.27 ^ 0.07
Equatorial 0.53 ^ 0.07 0.33 ^ 0.07 0.27 ^ 0.07 0.33 ^ 0.07
Apical 0.60 ^ 0.07 0.20 ^ 0.07 0.20 ^ 0.07 0.47 ^ 0.07
ECC
Basal 20.07 ^ 0.07 20.13 ^ 0.07 20.07 ^ 0.07 20.07 ^ 0.07
Equatorial 20.13 ^ 0.07 20.13 ^ 0.07 20.07 ^ 0.07 20.20 ^ 0.07
Apical 20.13 ^ 0.07 20.07 ^ 0.07 20.07 ^ 0.07 20.20 ^ 0.07
ELL
Basal 20.13 ^ 0.07 20.13 ^ 0.07 20.13 ^ 0.07 20.13 ^ 0.07
Equatorial 20.13 ^ 0.07 20.13 ^ 0.07 20.07 ^ 0.07 20.07 ^ 0.07
Apical 20.20 ^ 0.07 20.13 ^ 0.07 20.13 ^ 0.07 20.13 ^ 0.07
340 F. Dorri et al.

models were mostly evaluated with experimental results arrangement leads to distinct abnormalities in the
of strain using epi-cardial markers at a limited number of ventricular deformation pattern. The interpretation of
measurement sites (Guccione et al. 1995, Nash and Hunter such abnormalities can essentially be improved by
2000, Usyk et al. 2000). Yet, some of these models could applying mathematical approaches. Even when assuming
not represent the global deformation of the LV correctly: that the ventricular fibre architecture is optimal with
A low ejection fraction was reported in one case respect to normal cardiac performance (Rijcken et al.
(Guccione et al. 1995) (in some other models no ejection 1997, 1999), this is at best a thumb rule which holds for a
fraction is given), furthermore longitudinal elongation healthy heart, yet, in case of pathologies such as fibrosis,
instead of shortening was predicted (Bovendeerd et al. infarction or cardiomyopathy the fibre field must be
1992, Guccione et al. 1995). An accurate comparison of expected to deviate significantly from an optimal pattern.
the results obtained with such idealised models with our
calculations, which were performed on an unsymmetrical
geometry including non surface-parallel fibres and a more
References
advanced active stress tensor is not straightforward.
Nevertheless, as far as comparisons were possible, general T. Arts, P.C. Veenstra and R.S. Reneman, “Epicardial deformation and
agreement was observed. left ventricular wall mechanics during ejection in the dog”, Am.
The development of cross-fibre stresses was studied J. Physiol., 243, pp. H379–H390, 1982.
P.H. Bovendeerd, T. Arts, J.M. Huyghe, D.H. van Campen and R.S.
experimentally (Lin and Yin 1998). These measurements Reneman, “Dependence of local left ventricular wall mechanics on
along with other observations suggest that fibre, cross- myocardial fiber orientation: a model study”, J. Biomech., 25,
fibre and shear stresses develop simultaneously in the pp. 1129–1140, 1992.
P.H. Bovendeerd, J.M. Huyghe, T. Arts, D.H. van Campen and R.S.
myocardium and should be taken into account for the Reneman, “Influence of endocardial–epicardial crossover of muscle
modelling of systolic contraction (Mazhari and McCul- fibers on left ventricular wall mechanics”, J. Biomech., 27,
loch 1999, Usyk et al. 2000). Yet, the configuration of an pp. 941–951, 1994.
P.H. Bovendeerd, T. Arts, T. Delhaas, J.M. Huyghe, D.H. van Campen
active stress tensor depends on the reference coordinate and R.S. Reneman, “Regional wall mechanics in the ischemic left
system; in an appropriately defined local coordinate ventricle: numerical modeling and dog experiments”, Am. J. Physiol.,
system an active stress tensor can be implemented with 270, pp. H398–H410, 1996.
R.S. Chadwick, “Mechanics of the left ventricle”, Biophys. J., 39,
components corresponding to the local fibre field. In our pp. 279–288, 1982.
mathematical model we furthermore concluded that R.S. Chadwick, J. Ohayon and M. Lewkowicz, “Wall-thickness and
S0active ðS; SÞ should be negligible in comparison with the midwall-radius variations in ventricular mechanics”, Proc. Natl.
Acad. Sci. USA, 86, pp. 2996–2999, 1989.
(F, F) and (N, N) components because an active C.J. Chuong and Y.C. Fung, Frontiers in Biomechanics, G.W. Schmid-
compressive (S, S) component would counteract wall Schönbein, S.L.-Y. Woo and B.W. Zweifach, Eds., New York, NY:
thickening and longitudinal shortening, apart from that, Springer-Verlag, 1986, pp. 117 –129.
K.D. Costa, Y. Takayama, A.D. McCulloch and J.W. Covell, “Laminar
there is no experimental support for the existence of fiber architecture and three-dimensional systolic mechanics in canine
tensile (S, S) active components. ventricular myocardium”, Am. J. Physiol., 276, pp. H595–H607,
In this work we attempted to determine deformation 1999.
C.W. Cryer, H. Navidi-Kasmai, P.P. Lunkenheimer and K. Redmann,
patterns of the LV by inverting the modelling process, i.e. “Computation of the alignment of myocardial contractile pathways
to extrapolate stresses from deformations rather than using a magnetic tablet and an optical method”, Technol. Health
determining deformations from assumed fibre stresses. Care, 5, pp. 79 –93, 1997.
F. Dorri, “A finite element model of the human”, 2004, PhD Thesis, Swiss
Given the relation between the heart muscle’s oxygen Federal Institute of Technology Zurich.
consumption and wall stress, the clinical intention is to J.G. Dumesnil and R.M. Shoucri, “Quantitative relationships between left
evaluate the cardiac load and work performed by the heart. ventricular ejection and wall thickening and geometry”, J. Appl.
Physiol., 70, pp. 48 –54, 1991.
The ultimate goal of our work therefore consists of an H. Feneis, “Das Gefüge des herzmuskels bei systole und diastole”,
estimation of wall stresses from a known deformation Morphol. Jahrb., 89, pp. 371–406, 1943.
pattern. Our results indicate that this goal might J.M. Guccione and A.D. McCulloch, “Mechanics of active contraction in
cardiac muscle: Part I—constitutive relations for fiber stress that
successfully be reached; however, a more advanced describe deactivation”, J. Biomech. Eng., 115, pp. 72–81, 1993.
constitutive description of the heart muscle tissue is J.M. Guccione, A.D. McCulloch and L.K. Waldman, “Passive material
necessary. Nevertheless, our model included essential properties of intact ventricular myocardium determined from a
cylindrical model”, J. Biomech. Eng., 113, pp. 42–55, 1991.
anatomical details of a human heart, namely surface J.M. Guccione, K.D. Costa and A.D. McCulloch, “Finite element stress
irregularities and a realistic, albeit individual fibre analysis of left ventricular mechanics in the beating dog heart”,
orientation pattern. As expected, the analysis reveals that J. Biomech., 28, pp. 1167–1177, 1995.
J.M. Guccione, S.M. Moonly, P. Moustakidis, K.D. Costa, M.J. Moulton,
a realistic geometrical model along with a typical fibre M.B. Ratcliffe and M.K. Pasque, “Mechanism underlying mechanical
arrangement leads to considerable inhomogeneities in the dysfunction in the border zone of left ventricular aneurysm: a finite
systolic deformation pattern. This result holds indepen- element model study”, Ann. Thorac. Surg., 71, pp. 654–662, 2001a.
J.M. Guccione, S.M. Moonly, A.W. Wallace and M.B. Ratcliffe,
dent of further features influencing the constitutive “Residual stress produced by ventricular volume reduction surgery
behaviour of cardiac tissue. Our data support the has little effect on ventricular function and mechanics: a finite
assumption that accurate measurement of local systolic element model study”, J. Thorac. Cardiovasc. Surg., 122,
pp. 592–599, 2001b.
wall thickening, twisting and longitudinal shortening is of P.J. Hanley, A.A. Young, I.J. LeGrice, S.G. Edgar and D.S. Loiselle, “3-
diagnostic value in so far as any pathologic fibre Dimensional configuration of perimysial collagen fibres in rat cardiac
FE model of the human left ventricular systole 341

muscle at resting and extended sarcomere lengths”, J. Physiol., 517, V.P. Novak, F.C. Yin and J.D. Humphrey, “Regional mechanical
pp. 831–837, 1999. properties of passive myocardium”, J. Biomech., 27, pp. 403 –412,
K.B. Harrington, F. Rodriguez, A. Cheng, F. Langer, H. Ashikaga, G.T. 1994.
Daughters, J.C. Criscione, N.B. Ingels and D.C. Miller, “Direct F.E. Rademakers, W.J. Rogers, W.H. Guier, G.M. Hutchins, C.O. Siu,
measurement of transmural laminar architecture in the anterolateral M.L. Weisfeldt, J.L. Weiss and E.P. Shapiro, “Relation of regional
wall of the ovine left ventricle: new implications for wall thickening cross-fiber shortening to wall thickening in the intact heart. Three-
mechanics”, Am. J. Physiol. Heart Circ. Physiol., 288, dimensional strain analysis by NMR tagging”, Circulation, 89,
pp. H1324–H1330, 2005. pp. 1174–1182, 1994.
A. Horowitz, M. Perl, S. Sideman and E. Ritman, “Comprehensive model J. Rijcken, P.H. Bovendeerd, A.J. Schoofs, D.H. van Campen and T. Arts,
for the simulation of left ventricle mechanics. Part 2. Implementation “Optimization of cardiac fiber orientation for homogeneous fiber
and results analysis”, Med. Biol. Eng. Comput., 24, pp. 150– 156, strain at beginning of ejection”, J. Biomech., 30, pp. 1041– 1049,
1986. 1997.
W. Hort, “Makroskopische und mikrometrische Untersuchungen am J. Rijcken, P.H. Bovendeerd, A.J. Schoofs, D.H. van Campen and T. Arts,
Myokard verschieden stark gefüllter linker Kammern”, Virchows. “Optimization of cardiac fiber orientation for homogeneous fiber
Arch. Pathol. Anat. Physiol., 333, pp. 569– 581, 1960. strain during ejection”, Ann. Biomed. Eng., 27, pp. 289–297, 1999.
J.D. Humphrey, R.K. Strumpf and F.C. Yin, “Determination of a R.S. Rivlin and D.W. Saunders, “Large elastic deformations of isotropic
constitutive relation for passive myocardium: II. Parameter materials, VII experiments on the deformation of rubber”, Philos.
estimation”, J. Biomech. Eng., 112, pp. 340 –346, 1990a. Trans. R. soc., A243, pp. 251–288, 1951.
J.D. Humphrey, R.K. Strumpf and F.C. Yin, “Determination of a D.F. Rogers, p An q Introduction to NURBS with Historical
constitutive relation for passive myocardium: I. A new functional Perspective, San Francisco, CA: Morgan Kaufmann, 2001.
form”, J. Biomech. Eng., 112, pp. 333 –339, 1990b. H.N. Sabbah, M. Marzilli and P.D. Stein, “The relative role of
P.J. Hunter, P.M. Nielsen, B.H. Smaill, I.J. LeGrice and I.W. Hunter, “An subendocardium and subepicardium in left ventricular mechanics”,
anatomical heart model with applications to myocardial activation Am. J. Physiol., 240, pp. H920–H926, 1981.
and ventricular mechanics”, Crit. Rev. Biomed. Eng., 20, R.M. Shoucri, “Active and passive stresses in the myocardium”, Am.
pp. 403–426, 1992. J. Physiol. Heart Circ. Physiol., 279, pp. H2519–H2528, 2000.
P.J. Hunter, A.D. McCulloch and H.E. Keurs, “Modelling the mechanical A.J. Sinusas, X. Papademetris, R.T. Constable, D.P. Dione, M.D. Slade, P.
properties of cardiac muscle”, Prog. Biophys. Mol. Biol., 69, pp. Shi and J.S. Duncan, “Quantification of 3-D regional myocardial
289 –331, 1998. deformation: shape-based analysis of magnetic resonance images”,
J.M. Huyghe, T. Arts, D.H. van Campen and R.S. Reneman, “Porous Am. J. Physiol. Heart Circ. Physiol., 281, pp. H698–H714, 2001.
medium finite element model of the beating left ventricle”, Am. A.J.M. Spencer, Deformations of Fibre-Reinforced Materials, Oxford:
J. Physiol., 262, pp. H1256–H1267, 1992. Clarendon Press, 1972.
N.B. Ingels, “Myocardial fiber architecture and left ventricular function”, A.J.M. Spencer, Continuum Theory of the Mechanics of Fibre-Reinforced
Technol. Health Care, 5, pp. 45–52, 1997. Composites, Wien: Springer, 1984.
I.J. LeGrice, B.H. Smaill, L.Z. Chai, S.G. Edgar, J.B. Gavin and P.J. H.M. Spotnitz, W.D. Spotnitz, T.S. Cottrell, D. Spiro and E.H.
Hunter, “Laminar structure of the heart: ventricular myocyte Sonnenblick, “Cellular basis for volume related wall thickness
arrangement and connective tissue architecture in the dog”, Am. changes in the rat left ventricle”, J. Mol. Cell Cardiol., 6,
J. Physiol., 269, pp. H571 –H582, 1995. pp. 317–331, 1974.
I.J. LeGrice, P.J. Hunter and B.H. Smaill, “Laminar structure of the heart: C. Stevens, E. Remme, I. LeGrice and P. Hunter, “Ventricular mechanics
a mathematical model”, Am. J. Physiol., 272, pp. H2466–H2476, in diastole: material parameter sensitivity”, J. Biomech., 36,
1997. pp. 737–748, 2003.
D.H.S. Lin and F.C. Yin, “A multiaxial constitutive law for mammalian D.D. Streeter, JR and D.L. Basset, “An engineering analysis of
left ventricular myocardium in steady-state barium contracture or myocardial fibre orientation in pig’s left ventricle in systole”, Anat.
tetanus”, J. Biomech. Eng., 120, pp. 504–517, 1998. Rec., 155, pp. 503 –511, 1966.
P.P. Lunkenheimer, K. Redmann, K.-H. Dietl, C. Cryer, K.-D. Richter, D.D. Streeter, JR, D.P.P. Spotnitz, J.J. Ross and E.H. Sonnenblick, “Fiber
W.F. Whimster and P. Niederer, “The heart’s fibre alignment assessed orientation in the canine left ventricle during diastole and systole”,
by comparing two digitizing systems. Methodological investigation Circ. Res., 24, pp. 339–347, 1969.
into the inclination angle toward wall thickness”, Technol. Health M. Stuber, Quantification of the Human Heart Wall Motion by Ultra-fast
Care, 5, pp. 65–77, 1997. Magnetic Resonance Myocardial Tagging Techniques, Zurich: Swiss
P.P. Lunkenheimer, K. Redmann, J. Florek, U. Fassnacht, C.W. Cryer, F. Federal Institute of Technology, 1997.
Wübbeling, P. Niederer and R.H. Anderson, “The forces generated M. Stuber, M.B. Scheidegger, S.E. Fischer, E. Nagel, F. Steinemann,
within the musculature of the left ventricular wall”, Heart, 90, O.M. Hess and P. Boesiger, “Alternations in the local myocardial
pp. 200–207, 2004. motion pattern in patients suffering from pressure overload due to
G.A. MacGowan, E.P. Shapiro, H. Azhari, C.O. Siu, P.S. Hees, G.M. aortic stenosis”, Circulation, 100, pp. 361–368, 1999.
Hutchins, J.L. Weiss and F.E. Rademakers, “Noninvasive measure- F. Torrent-Guasp, W.F. Whimster and K. Redmann, “A silicone rubber
ment of shortening in the fiber and cross-fiber directions in the normal mould of the heart”, Technol. Health Care, 5, pp. 13 –20, 1997.
human left ventricle and in idiopathic dilated cardiomyopathy”, F. Torrent-Guasp, G.D. Buckberg, C. Clemente, J.L. Cox, H.C. Coghlan
Circulation, 96, pp. 535 –541, 1997. and M. Gharib, “The structure and function of the helical heart and its
F.P. Mall, “On the muscular architecture of the ventricles of the human buttress wrapping. I. The normal macroscopic structure of the heart”,
heart”, Am. J. Anat., 11, pp. 211– 266, 1911. Semin. Thorac. Cardiovasc. Surg., 13, pp. 301 –319, 2001.
R. Mazhari and A.D. McCulloch, Proceeding of ASME 4th Summer W.Y. Tseng, T.G. Reese, R.M. Weisskoff, T.J. Brady and V.J. Wedeen,
Bioengineering Conference, ASME, ed., 1999, pp. 43–44. “Myocardial fiber shortening in humans: initial results of MR
A.D. McCulloch, L.K. Waldman and J.M. Rogers, “Large-scale finite imaging”, Radiology, 216, pp. 128– 139, 2000.
element analysis of the beating heart”, Crit. Rev. Biomed. Eng., 20, T.P. Usyk, R. Mazhari and A.D. McCulloch, “Effect of laminar
pp. 427–449, 1992. orthotropic myofiber architecture on regional stress and strain in the
C.C. Moore, C.H. Lugo-Olivieri, E.R. McVeigh and E.A. Zerhouni, canine left ventricle”, J. Elasticity, 61, pp. 143 –164, 2000.
“Three-dimentional systolic strain patterns in the normal human left L.K. Waldman, Y.C. Fung and J.W. Covell, “Transmural myocardial
ventricle: characterization with tagged MR imaging”, Radiology, deformation in the canine left ventricle. Normal in vivo three-
214, pp. 453– 466, 2000. dimensional finite strains”, Circ. Res., 57, pp. 152–163, 1985.
P. Moustakidis, R. Pyo and R.P. Scheri, “Stress distribution analysis of an J.A. Weiss, “A constitutive model and finite element representation for
experimental model of the left ventricular aneurysm using magnetic transversely isotropic soft tissue”, PhD Thesis, The University of
resonance imaging and finite element modeling”, Circulation, 100, Utha. 1994.
p. 644, 1999. F.C. Yin, C.C. Chan and R.M. Judd, “Compressibility of perfused passive
M.P. Nash and P.J. Hunter, “Computational mechanics of the heart”, myocardium”, Am. J. Physiol. Heart Circ. Physiol., 40,
J. Elasticity, 61, pp. 113 –141, 2000. pp. H1864–H1870, 1996.
P.M. Nielsen, I.J. Le Grice, B.H. Smaill and P.J. Hunter, “Mathematical A.A. Young, I. LeGrice, M.A. Young and B.H. Smaill, “Extended
model of geometry and fibrous structure of the heart”, Am. J. Physiol., confocal microscopy of myocardial laminae and collagen network”,
260, pp. H1365–H1378, 1991. J. Microsc., 192, pp. 139–150, 1998.

You might also like