Precise Fluid-Solid Simulation of Human Left Ventricle along with
Aortic Valve during Systole
02+$00$'$/,021)$5(' A, MOHAMMAD MEHDI ALISHAHI A,1,
MARZIEH ALISHAHI A
a School of Mechanical Engineering,
Shiraz University,
Shiraz, 71936-16548, IRAN
Abstract: This paper presents an accurate blood flow model with tissue deformation of the human left
ventricle, including the aortic valve. A two-way fluid-solid Interaction (FSI) algorithm is employed to
simulate the performance of the human left ventricle during systole. The initial geometry of the left
ventricle is extracted from CT scan images of a healthy person. The simulation results produced the
systolic anterior motion of the Left Ventricle (LV) identical with the CT scan images at later times during
systole. Besides, the numerical results for left ventricular volume change, maximum blood velocity at the
aortic valve, and its maximum opening are in good agreement with physiological data. Although no clear
image of the aortic valve is apparent in CT images, the FSI simulation predicted the maximum opening of
the aortic valve to be 4.38 cm2 which is consistent with physiological observation on a healthy individual.
As an application of the above algorithm, a model of Hypertrophic Cardiomyopathy (HCM) or septal wall
thickening disease is constructed and studied during systole. This simulation provides an understanding of
heart performance under HCM conditions. According to the simulation outcomes, the mitral valve
approaches the septal wall under HCM due to the change in pressure gradient and the drag force on the
mitral valve. This blockage of the LV blood passage by the mitral valve results in stagnation pressure loss
and weaker hearth pumping power. Therefore, the maximum opening of the aortic valve, in this case, is
2.28 cm2, which is much lower than the physiological range, indicating the drastic effect of HCM on the
performance of the aortic valve and systolic performance.
Keywords: CFD, Systole, Left ventricle, Aortic valve, Fluid-Solid interaction, Mitral valve.
Received: April 15, 2021. Revised: January 12, 2022. Accepted: January 22, 2022. Published: March 3, 2022.
WSEAS TRANSACTIONS on FLUID MECHANICS
DOI: 10.37394/232013.2022.17.3
Mohammadali Monfared,
Mohammad Mehdi Alishahi,
Marzieh Alishahi
E-ISSN: 2224-347X
18
Volume 17, 2022
1 Introduction
Cardiovascular diseases are the most crucial
cause of death, about 17.9 million people, an
estimated 31% of all deaths worldwide.
Therefore, any attempt to provide a better
knowledge of how they affect the regular
operation of the heart is worthwhile. One of
the disciplines for such studies is the
biomechanical branch and specifically the
heart simulation. In the end, such the ability
to predict the malfunction of the diseased
heart would be invaluable for choosing a
cure strategy.
With the ongoing growth of computational
capacities, researchers are inclined to use
Computational Fluid Dynamics (CFD)
methods and especially Fluid-Solid
Interaction (FSI) to simulate blood flow in
different parts of the cardiovascular system
[1-3]. These methods are also employed to
investigate heart pumping function in a
simplified model of the Left Ventricle (LV)
along with the aortic valve using different
geometrical assumptions [4]. Watanabe et
al. [5] assumed a simplified geometry model
and investigated the ventricular volume
change due to different contractile-stimulus
using the FSI method. The computational
algorithm was based on the Eulerian-
Lagrangian finite element method, while the
dynamic meshing captured large
deformations. Moosavi et al. [6] constructed
an anatomical model of the LV and the
aortic sinus based on the real images of a
volunteer and carried out the numerical
simulations of blood flow in the model. The
computed results for aortic outflow were
compared with the data obtained by the
phase-contrast MRI, and an acceptable
agreement between the simulation results
and the physiological measurements was
reported. Lorenz et al. [7] presented a
cardiac model consisting of four cardiac
chambers, cardiovascular and coronary
arteries. This geometric heart model was
built based on data from 27 CT cores at the
end of the diastole. The model's accuracy
was measured based on the comparison with
different references.
There is also some research on more precise
aortic valve modeling. Kim et al. [8]
simulated the nonlinear structure of polymer
aortic valves with three leaflets using both
computational and experimental approaches.
They presented a structural model that can
predict the heart structure's transformation
with a maximum error of 10% during one
complete stage of heart contraction.
Moreover, another research carried out by
Merryman et al. [9] studied the function of
the heart valves by estimating at least 3 ×
109 cycles for one person's life. Focused on
the biocompatibility of heart valves, i.e.,
bio-solid along with the bio-fluid
characteristics, the physiological function of
the aortic valve was well-documented.
Trung Bao Le et al. [10] investigated a real-
shaped LV with a simple aortic valve model,
two sheets that open and close, using the FSI
method. In this modeling, the mitral valve
was assumed to be completely open during
diastole and completely closed during
systole. The results showed an asymmetry in
the closure of aortic valve leaflets, which did
not match the physiological results.
Considering the material models available
for valve’s leaflet tissue, Sturla et al. [11]
assumed the mechanical response of the
aortic valve to be linear, and therefore
elastic material was an appropriate model.
WSEAS TRANSACTIONS on FLUID MECHANICS
DOI: 10.37394/232013.2022.17.3
Mohammadali Monfared,
Mohammad Mehdi Alishahi,
Marzieh Alishahi
E-ISSN: 2224-347X
19
Volume 17, 2022
Employing the FSI method, it was shown
that the use of elastic material for valve
leaflets gives relatively good results
compared to experimental results. MC Hsu
et al. [12] developed a hyperelastic model of
a biological aortic valve placed at the
opening of the aorta. Two material models
were chosen for aortic valve tissue; Fung
and Venant-Kirchhoff models. It was
observed that the Venant-Kirchhoff model
produced relevant results. Using the FSI
method with the Venant-Kirchhoff model
for the aortic valve, the outcomes were
remarkably close to the measured in vivo
data.
The main objective of the current study is to
present and investigate a real-shaped model
of LV along with an aortic valve during
systole. The initial 3-dimensional geometry
is constructed based on CT scan images of a
healthy person to fulfill this purpose. Blood
is considered a Newtonian fluid, and the
hyperelastic Ogden model is chosen for
ventricular wall tissue and aortic valve
leaflets. The blood flow inside the LV and
ventricular wall tissue movement were
simulated using a two-way FSI algorithm,
and the obtained results are compared with
physiological data as a verification of the
modeling and the simulation. Thorough
apprehension of aortic valve plus LV
operation provide a more precise tool for
studying different diseases and could assist
heart valve pre-surgery evaluations. Along
this path, a model of septal wall thickening
disease, hypertrophic cardiomyopathy
(HCM), is also constructed and investigated
during the systole. HCM is a heart muscle
disease where the myocardium becomes
thickened, resulting in stiffer heart muscle.
If HCM occurs in the LV, it also will reduce
the ventricular volume up to 50% [13]. In
the case of HCM, the mitral valve leaflets
approach the septum wall and partially block
the blood flow into the aorta, and eventually,
the heart will lose its pumping ability. There
are two potential causes for such
phenomena; the change in pressure gradient
and drag force hypothesis [14] resulting in
total pressure loss in systolic performance.
In this paper, a simple form of HCM in LV
is modeled at the start of the systole, and the
resulting dislocation of mitral valve leaflets
and heart performance degradation is
investigated.
2 Physical Modeling
2-1 Left Ventricle Geometry
In this study, the geometry of the left
ventricle is extracted from CT scan images
of a healthy young male at the beginning of
systole. The CT scan images are taken by a
Philips Brilliance 64-slices CT scanner. The
images are captured in 21 phases, and each
phase contains 397 slices in three different
views. The first and last phases are identical.
The corresponding time and spatial
resolution are 38 ms and 0.42 mm,
respectively. The left ventricle geometry
was constructed from a CT angiography set
of images at the beginning of the systole
phase, using Mimics software. Figure 1
shows the CT images taken from one
volunteer at the beginning of the systole.
The red color represents the blood flow
region, and the violet color identifies the left
ventricle wall tissues.
WSEAS TRANSACTIONS on FLUID MECHANICS
DOI: 10.37394/232013.2022.17.3
Mohammadali Monfared,
Mohammad Mehdi Alishahi,
Marzieh Alishahi
E-ISSN: 2224-347X
20
Volume 17, 2022
The smoothed left ventricle geometries and
the ventricular volume, which is the blood region at the beginning of systole, are shown
in figure 2.
(a)
(b)
Figure 2 Left ventricle geometry (a) solid region (b) fluid region.
The HCM model is constructed based on the
geometry of reference [14] and is shown in
figure 3. It is noteworthy that the flexible
mitral valve is modeled for HCM study in
order to examine the septum and mitral
valve interactions.
(d)
Figure 1 Different views of left ventricle (a) Front view (b) Top view (c) Side view (d) 3D view.
Mitral inlet
Aortic outlet
(b)
WSEAS TRANSACTIONS on FLUID MECHANICS
DOI: 10.37394/232013.2022.17.3
Mohammadali Monfared,
Mohammad Mehdi Alishahi,
Marzieh Alishahi
E-ISSN: 2224-347X
21
Volume 17, 2022
Figure 3 Left ventricle geometry for HCM model (a) solid region (b) fluid region.
2-2 Aortic Valve Geometry
The aortic valve has a complex structure, but
due to the rapid opening and closure, proper
medical images of the valve could not be
extracted from CT images. Therefore, aortic
valve geometry is constructed with the aid
of CAD software, enclosed by boundaries of
CT scan images, and based on physiological
information. The dimensions of different
parts of the reconstructed aortic valve are
shown in figure 4.
(a)
(b)
(c)
Figure 4 Aortic valve model (a) and (b) front view (c) top view.
After reconstruction of the aortic valve, it
was integrated with the left ventricles in
CAD software. The left ventricular
geometry and added aortic valve are shown
in figure 5.
25mm
Sinuses of
Valsalva
15mm
41.9mm
Sinotubular
Junction
Leaflets
25.63m
m
22.63m
m
10.55mm
Aortic
leaflets
Mitral valve
Thickened
Septum
Mitral
(a)
(b)
Aorta
WSEAS TRANSACTIONS on FLUID MECHANICS
DOI: 10.37394/232013.2022.17.3
Mohammadali Monfared,
Mohammad Mehdi Alishahi,
Marzieh Alishahi
E-ISSN: 2224-347X
22
Volume 17, 2022
Figure 5 Left ventricle geometries with the aortic valve (a) solid domain (b) fluid domain.
2-3 Material Models
Blood is modeled as an isothermal,
incompressible and Newtonian viscous fluid
with density of 1060 Kg/m3 and a dynamic
viscosity of 0.00345 Pa.s [15-16]. Blood
flow is assumed to be laminar [6, 17].
During systole and diastole, the heart
muscles undergo large deformations;
therefore, the hyperelastic material model is
the best choice to simulate these behaviors.
For the sake of simplicity, the material of
the aortic valve and left ventricular wall is
assumed to be homogenous, isotropic, and
hyperelastic. This study uses the Ogden
model for the hyperelastic strain energy
function for the aortic valve and left
ventricular wall. According to Ogden model,
the strain energy is defined as;


 󰇛󰇜
Where and are time dependent
variables, is the material compressibility
parameter and is principal stretch. The
initial shear modulus, and initial bulk
modulus are;

The corresponding parameters of the stress-
strain relation were previously evaluated
using the uniaxial experiment data [18-19]
and shown in table 1. The left ventricular
tissue and aortic valve densities are,
respectively, 1050 and 1040 kg/m3 [20].
Table 1. Ogden model parameters.
i=3
i=2
i=1
modelOgden rder O
rd
3
-2.814E-06
1.49E-06
4.56E-06
(Pa-1)
7.61
7.62
7.59
(2)
(1)
Aortic valve
Aortic valve
(a)
(b)
81.89
81.88
81.88
(Pa)
WSEAS TRANSACTIONS on FLUID MECHANICS
DOI: 10.37394/232013.2022.17.3
Mohammadali Monfared,
Mohammad Mehdi Alishahi,
Marzieh Alishahi
E-ISSN: 2224-347X
23
Volume 17, 2022
3 Numerical Modeling
This study applies the iterative segregated
method to solve the structure and fluid
equations using ANSYS software
separately. The coupling of fluid and solid
domains is carried out based on the two-way
FSI method [21]. Accordingly, the latest
results computed from one part of the
coupled system are transmitted to the other
part, and these communications are
iteratively continued until the desired
convergence criterion is reached. The whole
simulation is carried out for a total physical
time of 0.3 s, covering the whole systolic
period.
3-1 Governing Equations
The turbulent flow field of blood in LV is
described by the 3D Navier-Stokes
equations;

󰇍
(3)
󰇍
 󰇛
󰇍
󰇜
󰇍

󰇍
(4)
Where
󰇍
is the velocity vector, p is the
pressure, and are the density and the
viscosity of blood, respectively. The
equations (1) and (2) are solved by a
transient, pressure based, finite volume
algorithm.
3-2 Boundary Conditions (BC)
During the systole of a healthy heart, the
mitral valve is closed to prevent backflow to
the atrium while the aortic valve is open.
Therefore, the mitral valve is assumed as the
wall boundary condition with no flow
through in the present study. At the aortic
outlet, two different types of BC are used to
compare and evaluate the simulation results
with scan images. The first boundary
condition is the velocity BC, and the other is
the pressure BC at the aortic valve. Two
outlet BCs are the time-dependent
physiological velocity and physiological
pressure described respectively by Sagawa
et al. [22] and Fielden et al. [23]. Other
researchers widely used these relations for
blood flow simulation inside the aorta.
Figure 6 shows the physiological charts for
average pressure and average velocity
during systole.
Figure 6 The physiological charts during systole; (a) Average pressure BC [22] (b) Average velocity BC [23].
30
60
90
120
0 0,1 0,2 0,3
Pressure(mm-Hg)
Time(s)
0
0,2
0,4
0,6
0 0,1 0,2 0,3
Velocity(m/s)
Time(S)
(b)
(a)
WSEAS TRANSACTIONS on FLUID MECHANICS
DOI: 10.37394/232013.2022.17.3
Mohammadali Monfared,
Mohammad Mehdi Alishahi,
Marzieh Alishahi
E-ISSN: 2224-347X
24
Volume 17, 2022
The inner surfaces of the ventricular wall
and the fluid surface in contact with this
wall are assigned as fluid-structure
interaction BC. The summary of the
boundary conditions is demonstrated in
figure 7. The same BCs are defined for the
HCM model.
The FSI boundary condition can be
described as;
(5)
 
(6)
In the above equations, and are fluid
and solid displacements, and and are
fluid and solid stress tensors, respectively.
(a)
(b)
Figure 7 The boundary conditions; (a) solid boundary conditions (b) fluid boundary conditions
3-3 Computational Grid
For FSI modeling, the computational grids
are generated for fluid and solid domains,
and grid independence analysis was carried
out separately. The unstructured grids are
constructed for both solid and fluid regions.
The fluid grid consists of prismatic and
tetrahedral elements. The prismatic elements
are built near the wall surface to capture the
flow details in this region. The discretized
equations are solved for
each cell of the
computational grid. Three different grids in
both domains are generated to ensure that
the obtained results are independent of the
computational grid. Two surfaces are
defined in the fluid domain, one near the
apex and another near the aortic valve, to
study the grid independence of fluid domain
solution (figure 8). Figure 9 shows the grid
independence results regarding area-
weighted velocity computed on the
described planes at time 0.12 s.
Outlet BC:
Pressure BC
Velocity BC
Wall BC
Fluid-Solid
Interaction BC
Fix support BC
(Aorta outlet)
Fix support BC
(Mitral opening)
Fluid-Solid
Interaction BC
Free BC
WSEAS TRANSACTIONS on FLUID MECHANICS
DOI: 10.37394/232013.2022.17.3
Mohammadali Monfared,
Mohammad Mehdi Alishahi,
Marzieh Alishahi
E-ISSN: 2224-347X
25
Volume 17, 2022
(a)
(b)
Figure 9 Mesh independency analysis based on average velocity changes on (a) plane 1 (b) plane 2.
Three grids with 143227, 299640, and
427667 elements are built to determine the
optimal spatial mesh resolution. According
to figure 9, the results of the three grids are
identical, which means that the obtained
solution is independent of the fluid grid.
Also, in two grids with more elements, the
results are closer to each other. The average
difference in velocity magnitudes between
the model with 299640 elements and the
model with 427667 elements is less than
1%. Therefore, the grid with 427667
elements is selected for further study. It
should be noted that at the end of the
systole, the speed on planes 1 and 2 is not
zero, which is due to the incomplete aortic
valve closure.
Grid independence analysis is also
performed for the solid domain. Therefore,
three different solid grids were generated
with 107699, 151689, and 226413 elements.
The results for the solid domain are
presented in Table 2 for different numbers
of elements for the LV wall tissue and aortic
valve.
Table 2. Mesh study of the solid domain for the physical model.
Maximum deformation
of aortic valve (mm)
Maximum heart muscle
deformation (mm)
Number of aortic valve
elements
Number of heart
muscle elements
5.88
20.612
41583
66161
6.05
20.625
63425
88264
6.08
20.628
115291
111122
0
0,2
0,4
0,6
0,8
1
0 0,1 0,2 0,3
Velocity(m/s)
Time(s)
mesh 143227 mesh 299640 mesh 427667
0
0,1
0,2
0 0,1 0,2 0,3
Velocity(m/s)
Time(s)
mesh 143227 mesh 299640 mesh 427667
Plane 1
Plane 2
Figure 8 Position of planes used for grid
independence analysis.
WSEAS TRANSACTIONS on FLUID MECHANICS
DOI: 10.37394/232013.2022.17.3
Mohammadali Monfared,
Mohammad Mehdi Alishahi,
Marzieh Alishahi
E-ISSN: 2224-347X
26
Volume 17, 2022
A similar grid independence analysis is also
carried out for the HCM model. Three grids
for the fluid domain are constructed; 80939,
190217, and 291154 elements. The results of
the grid with 190217 elements are
independent of grid size, and it is considered
for further simulations.
Moreover, three time-steps, 0.0002 s, 0.0005
s, and 0.001s, were chosen to study the time
step independence of the simulation
outcomes. It was observed that the
simulation results with the time-step 0.0002
s were appropriate.
4 Result and Discussions
4-1 Solid Domain
The time variation of different parameters
related to the aortic valve and myocardium
tissue are shown in figure 10. At each node,
the deformation is calculated in three
Cartesian directions, and the maximum
value of the resultant deformations in the
entire geometry is reported as the maximum
deformation at any given time (figure 10a
and 10c). Based on figure 10, the
deformation profile of the aortic valve is
almost similar to the output velocity profile
(figure 9a). The maximum deformation of
the aortic valve occurred at 0.12 s, at the
same time where the output velocity is
maximum (figure 9a). The following fact
can explain it; as the aortic valve continues
opening and its deformation increases to a
maximum, the output velocity increases too
and reaches a maximum without any time
lag. However, the aortic valve opening
happens in the numerical model due to the
velocity or pressure BC at the outlet. It is
noteworthy that the deformation of the aortic
valve gradually decreases after this
maximum (figure 10a) while the velocity
drop is steeper (figure 9a) due to relaxation
of the heart muscle that is clearly shown in
figure 10d, after t=0.12 sec. Note that the
early stage of systole, i.e., isovolumetric
compression that happens quite quickly, is
not precisely modeled here, although the
steep pressure rises before t=9.5 millisecond
(figure 6a) somehow takes this effect into
account.
(a)
(b)
0
2
4
6
0,0 0,1 0,2 0,3
Max Def.(mm)
Time(s)
0
1
2
3
4
0,0 0,1 0,2 0,3
Max Stress(KPa)
Time(s)
WSEAS TRANSACTIONS on FLUID MECHANICS
DOI: 10.37394/232013.2022.17.3
Mohammadali Monfared,
Mohammad Mehdi Alishahi,
Marzieh Alishahi
E-ISSN: 2224-347X
27
Volume 17, 2022
(c)
(d)
Figure 10 (a) Maximum deformation of the aortic valve (b) Maximum Tension in Aortic Valve Tissue (c) Maximum
deformation of myocardium tissue (d) Maximum tension in myocardium.
According to figure 10a, the deformation is
almost zero at the end of the systole in
accordance with zero velocity at the aortic
valve. Note that a closed valve position
might accompany zero outflow. However,
negative cell volume may occur in the grid
in the event of complete closure of the aortic
valve. The proper precautions are taken to
avoid this erroneous situation in the current
study. At the start of the systole period, this
problem is handled by initializing the flow
field with a negligible velocity.
The blood volume in the left ventricle for
the specific person during systole was
obtained by Mimics software. To valid the
numerical outcomes, they should be
compared with the measured data of the that
individual whose CT scan images are used
for this research. The for numerical
simulations and measured data for blood
volume changes in the left ventricle are
shown in figure 11. According to figure 11,
the ventricular volumes decrease
monotonically in accordance with
physiological predictions for both
simulations with different BCs. This is in
line with the increasing deformation of
myocardium tissue shown in figure 10c and
stress relaxation in figure 10d. Moreover,
the trends of the simulation curves in figure
11 show a good agreement with the
physiological data and the experimental data
of the specific case.
Figure 11 The outlet BC effect on the computed left ventricular volume change
Secondly, as shown in figure 11, there is a
slight difference between the computed
systolic volume of LV with velocity BC and
pressure BC. Meanwhile, the difference of
these two simulations with physiological
0
5
10
15
20
25
0,0 0,1 0,2 0,3
Max Def.(mm)
Time(s)
0
2
4
6
0,0 0,1 0,2 0,3
Max Stress(KPa)
Time(s)
70
90
110
130
0 0,05 0,1 0,15 0,2 0,25 0,3
Volume(ml)
Time
Left ventricle volume - Pressure boundary condition
Left ventricle volume - physiological data
Left ventricle volume - velocity boundary condition
WSEAS TRANSACTIONS on FLUID MECHANICS
DOI: 10.37394/232013.2022.17.3
Mohammadali Monfared,
Mohammad Mehdi Alishahi,
Marzieh Alishahi
E-ISSN: 2224-347X
28
Volume 17, 2022
data is more sensible, 10% maximum which
tends to be smaller at the end of systole.
However, the actual LV shape variation
during systole is different for these two
BC’s, explained later.
This is expectable since by defining a
specific time variation of velocity at a
constant sectional area such as the aorta, i.e.,
velocity BC, we define the ventricle volume
change; hence the relative agreement of
velocity BC and the physiological
predictions was seen. However, one expects
more realistic results for pressure BC due to
more freedom for flow field allowed by such
BC.
Figure 12 demonstrates the aortic valve
deformations during systole for simulations
with applying pressure BC. The CT scan
images from [24] are also added for
comparison. The overall agreement is
apparent.
In Figure 13, LV deformations during
systole with outlet pressure BC are shown,
which can be compared with the
corresponding CT scan images collected in
this study. Different views are provided in
order to have a better comparison. Initial LV
shapes in all views, at t=0 Sec., are exactly
the same as the CT images, and at later
times differences between numerical and CT
images show the simulation discrepancies.
According to the CT scan image of figure
13a, in addition to movements of the left
ventricle muscle (mainly upward) that
enhances blood pumping, a thickening of
myocardium muscle, especially in the lower
parts of the ventricle, is clearly seen. These
effects are mainly due to stimulant electrical
currents that are not modeled in this study.
The agreement between the simulation
results and the CT images in all views for
time t=0.12 s is good and worsens at later
time-steps. The overall agreement of the
images is acceptable, and the main
difference is the abnormal side contraction
of the simulation images that might be
caused by the absence of the vertical
movement and thickening of the LV tissue,
as mentioned above. Also, left ventricle
systolic anterior motion is depicted in
simulation results. Horizontal vectors
represent this motion left and right in figure
13. This difference in ventricle shape after
the start of systole perhaps shows the
importance of electrical response modeling
of the heart tissue, which is absent in the
present study.
Model (a)
CT Scan images
Figure 12 Comparison of deformation of the aortic
valve for the performed simulation and CT scan
images [24].
WSEAS TRANSACTIONS on FLUID MECHANICS
DOI: 10.37394/232013.2022.17.3
Mohammadali Monfared,
Mohammad Mehdi Alishahi,
Marzieh Alishahi
E-ISSN: 2224-347X
29
Volume 17, 2022
(a)
Simulation of the first
ventricle-pressure B.C
CT Scan images of first
ventricle
Simulation of the first
ventricle-pressure B.C
CT Scan images of first
ventricle
Time= 0 s
Time= 0.12 s
Time= 0.24 s
Time= 0.3 s
(b) Isometric view
Time= 0 s
Time= 0.12 s
Time= 0.24 s
Time= 0.3 s
(c) Up view
Time= 0 s
Time= 0.12 s
WSEAS TRANSACTIONS on FLUID MECHANICS
DOI: 10.37394/232013.2022.17.3
Mohammadali Monfared,
Mohammad Mehdi Alishahi,
Marzieh Alishahi
E-ISSN: 2224-347X
30
Volume 17, 2022
Time= 0.24 s
Time= 0.3 s
(d) Front view
Time= 0 s
Time= 0.12 s
Time= 0.24 s
Time= 0.3 s
Figure 13 Comparison of LV deformation based on simulation and CT scan images from different views.
4-2 Fluid Domain
Table 3 lists the computed hemodynamic
parameters of LV along with the related
physiological data [25-26]. All the computed
parameters lie in the normal ranges.
Table 3. Comparison of hemodynamic parameters of left ventricle with physiological values.
Present study-
Pressure B.C-
First ventricle
Present study-
velocity B.C-
First ventricle
Physiological range
End-systolic volume (ml)
76.485
73.102
58.1 ± 30.1
Stroke-volume (ml)
62.166
65.92
55-100
Cardiac output (l/min)
4.74
5.017
4-8
End-diastolic volume (ml)
138.621
138.621
134.2 ± 39.9
Ejection fraction %
44.85
47.42
58.1 ± 11.9
The maximum blood velocity at the outlet
for a healthy aortic valve is smaller than 2
m/s, and the maximum cross-sectional area
at the outlet of the aortic valve is 3.9 ± 1.2
cm2 [27]. Regarding the velocity at plane 1
in figure 9a and the area ratio of the
maximum opening of the aorta valve at
t=0.12 s to the outlet area, the maximum
WSEAS TRANSACTIONS on FLUID MECHANICS
DOI: 10.37394/232013.2022.17.3
Mohammadali Monfared,
Mohammad Mehdi Alishahi,
Marzieh Alishahi
E-ISSN: 2224-347X
31
Volume 17, 2022
velocity of 1.607 m/s occurs at the aorta
valve, which is in the normal physiological
range. As shown in figure 14a, the mean
velocity at the exit of the aortic valve is
similar to the physiological reference chart
[27]. The only difference is in the magnitude
of maximum velocity. Although based on
the criteria that the maximum velocity
should be less than 2 m/s [27], the
simulation results are acceptable. The
maximum velocity at the output of the
domain, which is the same as the ascending
aorta, is presented in figure 14b.
(a)
(b)
Figure 14 Velocity profiles (a) at the exit of the aortic valve (b) at the ascending aorta.
4-3 HCM Results
As mentioned earlier, under HCM condition
the mitral valve leaflets approach towards
the septum wall. As the leaflets move within
a certain vicinity of the wall, some negative
volume elements occur in the neighboring
region. The HCM simulations are carried
out for the beginning of the systole up until
0.3s to avoid this computational error. A
significant parameter is named “coaptation-
to-septal distance” and is defined as the
shortest distance between the coaptation
point at the end-systole and the
intraventricular septum. Figure 15 shows
this distance and the geometry of the left
ventricle with the two valves added.
Figure 15 The coaptation-to-septal distance region and the geometry used for HCM simulation
Ventricular volume change is plotted for all
the simulations in figure 16 to compare
HCM results with the healthy condition.
0
0,5
1
1,5
2
0 0,1 0,2 0,3
Velocity(m/s)
Time(s)
Velocity profile at the exit of the aortic valve [25]
Velocity profile at the exit of the aortic valve - current
simulation
0
0,2
0,4
0,6
0 0,1 0,2 0,3
Velocity(m/s)
Time(s)
Velocity profile at the entrance of the aorta tube [25]
The velocity profile at the entrance of the aortic tube -
current simulation
WSEAS TRANSACTIONS on FLUID MECHANICS
DOI: 10.37394/232013.2022.17.3
Mohammadali Monfared,
Mohammad Mehdi Alishahi,
Marzieh Alishahi
E-ISSN: 2224-347X
32
Volume 17, 2022
Figure 16 The left ventricular volume change for HCM condition.
Considering the whole cardiac cycle, the
cardiac output for the HCM condition would
be 2.123 lit/min, which is much less than the
physiological value under healthy condition,
i.e., 4-8 lit/min. Figure 17 shows the 3D
streamlines for laminar flow under HCM
condition.
t=0.01s
t=0.04s
t=0.08s
t=0.12s
Figure 17 3D streamlines for HCM condition
In order to study the interaction of the
septum wall and mitral valve for the HCM
case, figure 17 is presented. According to
figure 17, the streamlines should rotate to
pass the mitral valve and through the narrow
50
70
90
110
130
0 0,05 0,1 0,15 0,2 0,25 0,3
Volume(ml)
Time(S)
Healthy Left Ventricle
Hypertropic Cardiomyopathy
Physiologic Volume
WSEAS TRANSACTIONS on FLUID MECHANICS
DOI: 10.37394/232013.2022.17.3
Mohammadali Monfared,
Mohammad Mehdi Alishahi,
Marzieh Alishahi
E-ISSN: 2224-347X
33
Volume 17, 2022
region between the mitral valve leaflets and
the septum wall (coaptation-to-septal
distance). This kind of flow exerts a drag
force on the mitral valve, which will make
the anterior leaflet move towards the septum
wall during the systolic anterior motion.
Based on the figure 18 and compared with a
healthy heart, the reduction in the cross-
section area and the decrease of the local
pressure along with the presence of the drag
force on the posterior leaflet and the lower
pressure on the anterior leaflet of a flexible
valve bring about more blockage of the LV
blood passage by the mitral valve.
Consequently, the flexible mitral valve
blockage results in stagnation pressure loss
and weaker heart-pumping performance.
Therefore, low local pressure and lower
stagnation pressure after the mitral valve
produce an adverse pressure gradient toward
the aorta and may even cause reverse flow.
This reverse flow is detected in simulation
results at t = 0.12s.
For HCM simulation, the maximum area of
the aortic valve is 2.228 cm2 and occurs at t
= 0.1s, while for healthy condition, the
maximum area of the aortic valve occurs at
the same time and is equal to 3.9±1.2 cm2.
Moreover, the maximum velocity at the
aorta inlet is 1.27 m/s and 0.414 m/s for
healthy and HCM conditions, respectively.
The comparison shows that under the HCM
condition, the blood velocity at the aorta's
inlet and the aortic valve's maximum
openness lie outside the physiological
ranges.
t=0.02s
t=0.04s
t=0.06s
WSEAS TRANSACTIONS on FLUID MECHANICS
DOI: 10.37394/232013.2022.17.3
Mohammadali Monfared,
Mohammad Mehdi Alishahi,
Marzieh Alishahi
E-ISSN: 2224-347X
34
Volume 17, 2022
t=0.1s
t=0.14s
t=0.16
t=0.3s
Figure 18 pressure contour around the coaptation-to-septal distance region for HCM condition
As shown in figure 18, the blood flow
initially moves to the aortic valve. It is
observed that during systole, the blood flow
behind the posterior leaflet of the mitral
valve moves it to the left ventricular wall
(septum). Streamline vectors in this form are
visible behind the posterior leaflet of the
mitral valve. The drag force on the mitral
valve causes it to move towards the septum
and reduces the distance between the
ventricular wall and the anterior leaflet of
the mitral valve. This distance is called
coaptation-to-septal distance. Another
prominent factor seen in the images of
figure 18 is the pressure gradient around the
mitral valve. Before the 0.12 second, as the
leaflets approach the septum, the pressure
after the mitral valve decreases. This
reduction in pressure after the mitral valve is
another factor that causes the anterior leaflet
of the mitral valve to move toward the
septum. The coaptation-to-septal distance is
reduced to the point that after the time 0.12 s
and the creation of stagnation pressure, the
flow begins to rotate after the mitral valve,
and the pressure gradient is reversed. At this
time, a reverse flow occurs.
5 Conclusion
The precise and correct simulation of the
human heart can be beneficial to estimate
the heart performance and cardiac output
under different in vivo conditions. In this
study, real CT scan images are used to
construct an actual shaped model of human
LV under the healthy and HCM disease
conditions. Using the FSI method, the blood
flow in LV and ventricular tissue and aortic
valve deformations are simulated during the
WSEAS TRANSACTIONS on FLUID MECHANICS
DOI: 10.37394/232013.2022.17.3
Mohammadali Monfared,
Mohammad Mehdi Alishahi,
Marzieh Alishahi
E-ISSN: 2224-347X
35
Volume 17, 2022
systole period of the cardiac cycle for two
models; the healthy LV and HCM model.
The 3D grid is built for physical models, and
the discretized equations are solved for both
solid and fluid domains. The grid
independence analysis is performed for both
models. Also, the time independence of
numerical outcomes is investigated to ensure
obtained results' accuracy. Two simulations
with different BCs at the aortic outlet are
carried out for the healthy model. The
velocity and pressure outlet BCs are defined
based on the physiological charts. The
results for left ventricular volume change for
these simulations are approximately the
same and in accordance with available
physiological data of LV volume change
during systole. While these two models
produce similar approximate flow and solid
results for systole, another study for the
diastole phase of the LV showed different
results for the velocity (imprecise) and
pressure BC (precise) models. The
explanation would be the strong effect of the
heart electrical current on the deformations,
which is presented during systole and not
modeled in the present study, and the
absence of this effect in the diastole phase.
For the healthy model, numerical results
capture the systolic anterior motion of LV
identical to the CT scan images. The blood
streamlines in LV display the local pressure
variation near the aortic valve leaflets during
the anterior motion. The results for the solid
domain show the good accordance between
FSI results for LV shape deformation and
CT scan images. Besides, the FSI results for
the aortic valve leaflets deformation are
analogous with available CT scan images of
this valve during systole. The maximum
deformation of the aortic valve occurred at
the time of 0.12 s; simultaneously, the
output velocity is maximum, and the aortic
leaflet experiences the maximal opening.
The maximum opening surface of the aortic
valve is 4.38 cm2 which is consistent with
physiological observation on a healthy
individual.
For the HCM model, the maximum opening
of the aortic valve is 2.228 cm2, which is not
in the physiological range, indicating the
drastic effect of HCM on the performance of
the aortic valve. Moreover, the ventricular
volume change decreases significantly from
healthy to HCM condition. Compatible with
the pathological nature of the HCM, under
the severe malady, the LV loses its pumping
ability to send blood into the aorta.
Accordingly, the maximum blood velocity at
the aortic outlet for HCM condition is 0.414
m/s while this velocity is 1.27 m/s for the
healthy condition. The comparison indicates
that under the HCM disorder, the blood
velocity at the aorta's inlet and the aortic
valve's maximum openness lie outside the
physiological spans.
In conclusion, the FSI simulation can
undoubtedly lead to feasible results for LV
simulation when relevant physical and
numerical details are considered. More
complete heart models considering heart
electrical current and subsequent tissue
deformation would eventually provide the
necessary numerical tool for practical
applications.
WSEAS TRANSACTIONS on FLUID MECHANICS
DOI: 10.37394/232013.2022.17.3
Mohammadali Monfared,
Mohammad Mehdi Alishahi,
Marzieh Alishahi
E-ISSN: 2224-347X
36
Volume 17, 2022
References
[1]
M. Alishahi, M. Alishahi and H. Emdad,
"Numerical simulation of blood flow in a
flexible stenosed abdominal real aorta," Scientia
Iranica, vol. 18, no. 6, pp. 1297-1305, 2011.
[2]
S. Numata, K. Itatani, K. Kanda, K. Doi, S.
Yamazaki, K. Morimoto, K. Manabe, K.
Ikemoto and H. Yaku, "Blood flow analysis of
the aortic arch using computational fluid
dynamics," European Journal of Cardio-
Thoracic Surgery , vol. 49, no. 6, pp. 1578-
1585, 2016.
[3]
F. Ismail, M. Shukri Zakaria, M. Tamagawa, A.
Fazli Abdul Azi, S. Wiriadidjaya, A. Azrif Basri
and K. Arifin Ahmad, "Computational fluid
dynamics study of blood flow in aorta using
OpenFOAM," Journal of Advanced Research in
Fluid Mechanics and Thermal Sciences, vol. 34,
no. 1, pp. 81-89, 2018.
[4]
T. B. Le and F. Sotiropoulos, "Fluidstructure
interaction of an aortic heart valve prosthesis
driven by an animated anatomic left ventricle,"
Journal of computational physics, vol. 244, pp.
41-62, 2013.
[5]
H. Watanabe, S. Sugiura, H. Kafuku and T.
Hisada, "Multiphysics simulation of left
ventricular filling dynamics using fluid-structure
interaction finite element method," Biophysical
journal, vol. 87, no. 3, pp. 2074-2085, 2004.
[6]
M.-H. Moosavi, N. Fatouraee, H. Katoozian, A.
Pashaei, O. Camara and A. F. Frangi,
"Numerical simulation of blood flow in the left
ventricle and aortic sinus using magnetic
resonance imaging and computational fluid
dynamics," Computer methods in biomechanics
and biomedical engineering, vol. 17, no. 7, pp.
740-749, 2014.
[7]
C. Lorenz and J. von Berg, "A comprehensive
shape model of the heart," Medical image
analysis, vol. 10, no. 4, pp. 657-670, 2006.
[8]
H. S. Kim, "Nonlinear multi-scale anisotropic
material and structural models for prosthetic and
native aortic heart valves," Georgia Institute of
Technology, Georgia, 2009.
[9]
M. S. Sacks, W. D. Merryman and D. E.
Schmidt, "On the biomechanics of heart valve
function," Journal of biomechanics, vol. 42, no.
12, pp. 1804-1824, 2009.
[10]
T. B. Le and F. Sotiropoulos, "Fluidstructure
interaction of an aortic heart valve prosthesis
driven by an animated anatomic left ventricle,"
Journal of computational physics, vol. 244, pp.
41-62, 2013.
[11]
F. Sturla, E. Votta, M. Stevanella, C. A. Conti
and A. Redaelli, "Impact of modeling fluid
structure interaction in the computational
analysis of aortic root biomechanics," Medical
engineering & physics, vol. 35, no. 12, pp. 1721-
1730, 2013.
[12]
M.-C. Hsu, D. Kamensky, Y. Bazilevs, M. S.
Sacks and T. J. Hughes, "Fluidstructure
interaction analysis of bioprosthetic heart
valves: significance of arterial wall
deformation," Computational mechanics, vol.
54, no. 4, pp. 1055-1071, 2014.
[13]
E. Braunwald, C. E. Seidman and U. Sigwart,
"Contemporary evaluation and management of
hypertrophic cardiomyopathy," Circulation, vol.
106, no. 11, pp. 1312-1316, 2002.
[14]
A. Moghaddaszade-Kermani, P. Oshkai and A.
Suleman, "Fluid-Structure Interaction
Simulation of Blood Flow Inside a Diseased
Left Ventricle With Obstructive Hypertrophic
Cardiomyopathy in Early Systole," In Fluids
Engineering Division Summer Meeting, vol.
43727, pp. 1799-1807, 2009.
[15]
W. Sun, K. Li and E. Sirois, "Simulated
elliptical bioprosthetic valve deformation:
implications for asymmetric transcatheter valve
deployment," Journal of biomechanics, vol. 43,
no. 16, pp. 3085-3090, 2010.
[16]
J. V. Soulis, G. D. Giannoglou, Y. S.
Chatzizisis, K. V. Seralidou, G. E. Parcharidis
and G. E. Louridas, "Non-Newtonian models for
molecular viscosity and wall shear stress in a 3D
reconstructed human left coronary artery,"
Medical engineering & physics 30, no. 1 (): .,
vol. 30, no. 1, pp. 9-19, 2008.
[17]
Y.-c. Fung, Biomechanics: circulation, Springer
Science & Business Media, 2013.
WSEAS TRANSACTIONS on FLUID MECHANICS
DOI: 10.37394/232013.2022.17.3
Mohammadali Monfared,
Mohammad Mehdi Alishahi,
Marzieh Alishahi
E-ISSN: 2224-347X
37
Volume 17, 2022
[18]
V. Prot, B. Skallerud, G. Sommer and G. A.
Holzapfel, "On modelling and analysis of
healthy and pathological human mitral valves:
two case studies," Journal of the mechanical
behavior of biomedical materials, vol. 3, no. 2,
pp. 167-177, 2010.
[19]
C. Yang, D. Tang, I. Haber, T. Geva and J.
Pedro, "In vivo MRI-based 3D FSI RV/LV
models for human right ventricle and patch
design for potential computer-aided surgery
optimization," Computers & structures, vol. 85,
no. 11, pp. 988-997, 2007.
[20]
A. M. Pouch, C. Xu, P. A. Yushkevich, A. S.
Jassar, M. Vergnat, J. H. Gorman, R. C.
Gorman, C. M. Sehgal and B. M. Jackson,
"Semi-automated mitral valve morphometry and
computational stress analysis 3D ultrasound,"
Journal of biomechanics , vol. 45, no. 5, pp.
903-907, 2012.
[21]
M. S. Arefin and Y. S. Morsi, "Fluid structure
interaction (FSI) simulation of the left ventricle
(LV) during the early filling wave (E-wave),
diastasis and atrial contraction wave (A-wave),"
Australasian physical & engineering sciences in
medicine, vol. 37, no. 2, pp. 413-423, 2014.
[22]
K. Sagawa, "The ventricular pressure-volume
diagram revisited," Circulation Research, vol.
43, no. 5, pp. 677-687, 1978.
[23]
S. W. Fielden, B. K. Fornwalt, M.
JeroschHerold, R. L. Eisner, A. E. Stillman
and J. N. Oshinski, "A new method for the
determination of aortic pulse wave velocity
using crosscorrelation on 2D PCMR velocity
data," Journal of Magnetic Resonance Imaging,
vol. 27, no. 6, pp. 1382-1387, 2008.
[24]
R. Haj-Ali, L. P. Dasi, H.-S. Kim, J. Choi, H. W.
Leo and A. P. Yoganathan, "Structural
simulations of prosthetic tri-leaflet aortic heart
valves," Journal of biomechanics, vol. 41, no. 7,
pp. 1510-1519, 2008.
[25]
T. Schlosser, K. Pagonidis, C. U. Herborn, P.
Hunold, K.-U. Waltering, T. C. Lauenstein and
J. Barkhausen, "Assessment of left ventricular
parameters using 16-MDCT and new software
for endocardial and epicardial border
delineation," American Journal of
Roentgenology, vol. 184, no. 3, pp. 765-773,
[26]
P. B. Raven, D. H. Wasserman, W. G. Squires
and T. D. Murray, Exercise Physiology, Nelson
Education, 2012.
[27]
R. Halevi, A. Hamdan, G. Marom, K. Lavon, S.
Ben-Zekry, E. Raanani, D. Bluestein and R. Haj-
Ali, "Fluidstructure interaction modeling of
calcific aortic valve disease using patient-
specific three-dimensional calcification scans,"
Medical & biological engineering & computing,
vol. 54, no. 11, pp. 1683-1694, 2016.
[28]
D. Chen, M. Müller-Eschner, H. von Tengg-
Kobligk, D. Barber, D. Böckler, R. Hose and Y.
Ventikos, "A patient-specific study of type-B
aortic dissection: evaluation of true-false lumen
blood exchange," Biomedical engineering
online, vol. 12, no. 1, pp. 1-16, 2013.
Contribution of individual authors
to the creation of a scientific article
Mohammad Monfared carried out the
simulation and writing down the manuscript.
Prof. Mohammad Mehdi Alishahi was the
academic supervisor who guided the
research.
Marzieh Alishahi wrote the final manuscript
and was responsible with editing process.
Sources of funding for research
presented in a scientific article or
scientific article itself
There is no specific source of funding for
this research.
2005.
WSEAS TRANSACTIONS on FLUID MECHANICS
DOI: 10.37394/232013.2022.17.3
Mohammadali Monfared,
Mohammad Mehdi Alishahi,
Marzieh Alishahi
E-ISSN: 2224-347X
38
Volume 17, 2022