You are on page 1of 56

BACHELOR’ S THESIS / BIOMEDICAL ENGINEERING 2023

Device-related thrombus in transcatheter


aortic valve implantation: a fluid-structure
interaction study

Manal Barrouhou Berrouhou


Device-related thrombus in transcatheter
aortic valve implantation: a fluid-structure
interaction study

Manal Barrouhou Berrouhou

Bachelor’s Thesis UPF 2022/2023

Thesis Supervisor(s):

Dr. Andy Luis Olivares Miyares

Carlos Albors Lucas


Acknowledgments
Firstly, I would like to acknowledge this thesis to my supervisors, Carlos Albors and
Andy Luis Olivares, for giving me the opportunity to develop this project and for
their continuous support throughout the entire process. I would also like to express
my sincere gratitude to Jordi Mill and Oscar Camara for their consistent encour-
agement and support, both academically and emotionally, during challenging times.
Their guidance and motivation proved invaluable in overcoming obstacles. Finally,
special mention to my colleague Laura Escot for being by my side throughout the
development of this project.

Last, but not least, I want to express my profound gratitude to my family and
friends for their encouragement and support throughout this journey.
Abstract
Transcatheter aortic valve implantation (TAVI) is a minimally invasive heart pro-
cedure that has emerged as a safe and effective treatment for patients with severe
symptomatic aortic stenosis (AS). This is a common heart pathology consisting of
the aortic valve’s narrowing leading to an obstruction of the left ventricular outflow
tract.

Recently, TAVI has been approved for all AS risk groups given the low compli-
cation rates reported. However, there are many post-TAVI complications such as
device-related thrombus (DRT). This is a multi-factorial problem that lacks com-
prehensive understanding and effective treatment. In this context, numerical simu-
lations have emerged as alternative tools to assess and predict thrombus formation.
Therefore, the objective of this study was to present a computational workflow for
modelling patient-specific TAVI under a fluid-structure interaction (FSI) approach
to study the risk of developing DRT post-TAVI.

Clinical data of one patient who underwent TAVI was provided by Hospital de
la Santa Creu i Sant Pau (Barcelona, Spain). The proposed workflow consisted of
the following steps: (1) acquiring patient-specific data; (2) developing the model
including patient-specific geometry and TAVI prosthesis; (3) setting up the FSI
simulation; (4) validating the results obtained; and (5) analyzing haemodynamic
thrombus indices to assess the risk of DRT after TAVI.

Sensitivity analyses were conducted to evaluate parameter influence on the FSI


simulation setup. This study lead to relevant conclusions to set the basis for future
work regarding the study of patient-specific TAVI modelling under an FSI approach
to address DRT.

Keywords
Transcatheter aortic valve implantation (TAVI), device-related thrombus (DRT),
fluid-structure interaction (FSI).
Preface
Despite the increasing use of transcatheter aortic valve implantation (TAVI) in clin-
ical practice over the past two decades, the factors influencing the risk of thrombus
formation following TAVI remain unclear. Current tools for assessing post-TAVI
thrombus formation lack detailed hemodynamic information. In this regard, in sil-
ico fluid-structure interaction (FSI) simulations offer a powerful means to compre-
hend the mechanical and hemodynamic interactions post-implantation and predict
outcomes such as device-related thrombosis (DRT). However, there is a need for
consensus in computational studies regarding the prediction of thrombosis risk after
TAVI. Therefore, this project aims to propose a workflow based on fluid-structure
interaction analysis that incorporates previous studies to evaluate the risk of DRT
after TAVI.

This study has been conducted in collaboration with cardiologist Abdel Hakim
Moustafa and biomedical engineers César Acebes Pinilla and Marián Iglesias Blanco
from Hospital de la Santa Creu i Sant Pau (Barcelona, Spain). Their contributions
include the assessment of the clinical application, understanding of the TAVI pro-
cedure and providing the necessary patient-specific data.
Index
1 Introduction 1
1.1 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 State of the art . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.2.1 Device-realted thrombus . . . . . . . . . . . . . . . . . . . . . 4
1.2.2 Computational models . . . . . . . . . . . . . . . . . . . . . . 6
1.3 Objectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

2 Methods 9
2.1 Clinical data acquisition . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.2 Anatomical 3D model geometry . . . . . . . . . . . . . . . . . . . . . 9
2.2.1 Patient-specific 3D model . . . . . . . . . . . . . . . . . . . . 9
2.2.2 TAVI prosthesis model . . . . . . . . . . . . . . . . . . . . . . 10
2.2.3 Simplified 3D model . . . . . . . . . . . . . . . . . . . . . . . 10
2.3 Solid domain model . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.3.1 Conversion of STL file to CAD file format . . . . . . . . . . . 11
2.4 Fluid domain model . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.5 Boundary conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.6 FSI simulation setup . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.6.1 Solid structure analysis . . . . . . . . . . . . . . . . . . . . . . 13
2.6.2 Fluid dynamics analysis . . . . . . . . . . . . . . . . . . . . . 14
2.6.3 Fluid-structure interaction coupling . . . . . . . . . . . . . . . 14
2.7 Sensitivity analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.7.1 Mesh convergence study . . . . . . . . . . . . . . . . . . . . . 17
2.7.2 Mechanical properties . . . . . . . . . . . . . . . . . . . . . . 17
2.7.3 Velocity analysis . . . . . . . . . . . . . . . . . . . . . . . . . 18
2.7.4 Under-relaxation factor analysis . . . . . . . . . . . . . . . . . 18
2.8 Mechanical and haemodynamic indices . . . . . . . . . . . . . . . . . 18

3 Results 20
3.1 Sensitivity analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
3.1.1 Mesh convergence study . . . . . . . . . . . . . . . . . . . . . 20
3.1.2 Mechanical properties . . . . . . . . . . . . . . . . . . . . . . 20
3.1.3 Velocity analysis . . . . . . . . . . . . . . . . . . . . . . . . . 22
3.1.4 Under-relaxation factor analysis . . . . . . . . . . . . . . . . . 22
3.2 FSI simulation results . . . . . . . . . . . . . . . . . . . . . . . . . . 23
3.2.1 Mechanical and haemodynamic indices . . . . . . . . . . . . . 23

4 Discussion 26
4.1 Limitations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
4.2 Future work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
4.3 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30

Bibliography 31
A Appendix 36
A.1 TAVI prosthesis CAD . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
A.2 Boundary conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
A.3 Mesh convergence study . . . . . . . . . . . . . . . . . . . . . . . . . 38
A.4 Ogden model results . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
A.5 Von Misses stress and shear stress . . . . . . . . . . . . . . . . . . . . 40
List of Figures
1 Transcatheter aortic valve implantation (TAVI) procedure example. A new
aortic valve is placed over the top of the old one through a cardiac catheter.
Retrieved from [1]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
2 Overview of families of bioprostheses for TAVI. The more transparent tran-
scatheter heart valves represent no longer available bioprostheses. Re-
trieved from [5]. TAVI = transcatheter aortic valve implantation. . . . . . 2
3 Clinical images. HeartNavigator software being used for TAVI procedure
planning. Retrieved from [9]. TAVI = transcatheter aortic valve implan-
tation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
4 Predictive model for thrombus formation after TAVI in the neo-sinus. Re-
trieved from [22]. V = velocity; TAVI = transcatheter aortic valve implan-
tation; TAV = transcatheter aortic valves. . . . . . . . . . . . . . . . . . 5
5 Cross-sectional view of detailed flow patterns for the TAVI Evolut R model.
The colour indicates the magnitude of the velocity through the aortic root,
with cooler colours indicating lower velocities and warmer colours indicat-
ing higher velocities. The time increment between frames is 0.1s. Retrieved
from [31]. U = velocity. . . . . . . . . . . . . . . . . . . . . . . . . . . 7
6 Proposed modelling workflow established for TAVI simulation using FSI
approach with a patient-specific geometry. TAVI = transcatheter aortic
valve implantation; FSI = fluid-structure interaction; CT = computed to-
mography; CAD = computer-aided design. . . . . . . . . . . . . . . . . . 9
7 (a) TAVI prosthesis metallic frame cell retrieved from [46], (b) sketch frame
reconstruction in SolidWorks from provided cell design in (a), (c) obtained
metallic frame cell after reconstruction. TAVI = transcatheter aortic valve
implantation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
8 Geometries used for the development of the simplified model, (a) TAVI
leaflets geometry recovered from [28], (b) representation of the aorta (tubu-
lar structure), (c) obtained simplified model for simulation. TAVI = tran-
scatheter aortic valve implantation. . . . . . . . . . . . . . . . . . . . . 11
9 Conversion from STL to CAD. (a) STL file of reference to which several
manual corrections were applied in Meshmixer. (b) Converted STL to CAD
file using ANSYS SpaceClaim 22.1 tools of correction and conversion. STL
= stereo-lithography; CAD = computer-aided design. . . . . . . . . . . . 12
10 Obtained CAD files. (a) Solid domain geometry. (b) Fluid domain ge-
ometry. (c) Both fluid and solid geometries where the fluid geometry is
enclosed within the solid geometry. CAD = computer-aided design. . . . . 12
11 (a) One-way and (b) two-way coupled FSI approaches. Retrieved from [52].
FSI = fluid-structure interaction. . . . . . . . . . . . . . . . . . . . . . . 15
12 Two-way FSI coupling between Transient Structural and Fluid Flow in
ANSYS Workbench 22.1. FSI = fluid-structure interaction. . . . . . . . . 16
13 FSI-derived blood flow velocity streamlines at four instances. The first row
(a-d) shows the top view of the leaflets opening and closing kinematics at
the corresponding instances. The second row shows the cross-sectional view
of detailed flow patterns. The colour indicates the magnitude of the velocity
through the aortic root at the centre plane of the valve, with cooler colours
indicating lower velocities and warmer colours indicating higher velocities.
Re-circulations at low velocities are marked down with a black circle. FSI
= fluid-structure interaction; t = time. . . . . . . . . . . . . . . . . . . . 24
14 Leaflet kinematics and FSI-derived total displacement (mm) at six in-
stances during leaflet opening and closing. FSI = fluid-structure interaction. 24
15 Leaflet kinematics and FSI-derived maximum principal stress (MPa) at six
instances during leaflet opening and closing. FSI = fluid-structure interaction. 25
A1 Metallic frame component designed in SolidWorks from literature measure-
ments and drawings [45, 56]. . . . . . . . . . . . . . . . . . . . . . . . . 36
A2 Metallic frame designed in SolidWorks inspired on TAVI frames design. . . 36
A3 (a) LVOT Doppler echocardiography of the patient, and (b) the velocity
curve segmented and post-processed to be adjusted and synchronized to the
patient-specific beat (0.882s per beat). LVOT = left ventricular outflow
tract. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
A4 Solid domain mesh with element sizes ranging from 1 - 4 mm, displaying
surface discontinuities in the leaflets, causing a deviation from the original
CAD geometry. CAD = computer-aided design. . . . . . . . . . . . . . . 38
A5 FSI-derived blood flow velocity streamlines at four instances. The top row
(a-d) shows the leaflets opening and closing kinematics at the corresponding
instances. FSI = fluid-structure interaction. . . . . . . . . . . . . . . . . 39
A6 Leaflet kinematics FSI-derived von Mises stress (MPa) at six instances
during leaflet opening and closing. FSI = fluid-structure interaction. . . . 40
A7 Leaflet kinematics FSI-derived shear stress (MPa) at six instances during
leaflet opening and closing. FSI = fluid-structure interaction. . . . . . . . 40
List of Tables
1 Review of boundary conditions and different modelling choices for TAVI
simulations reported in the literature. TAVI = transcatheter aortic valve
implantation; FSI = fluid-structure interaction; lit = literature; vels =
velocities; press = pressures; PS = patient-specific; wind = Windkessel;
conv = convective; BCs = boundary conditions. . . . . . . . . . . . . . . 6
2 Glutaraldehyde-treated bovine pericardium material properties. Retrieved
from [24]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
3 Review of mesh converge analysis performed over the solid and fluid do-
mains in ANSYS Meshing 22.1. . . . . . . . . . . . . . . . . . . . . . . 20
4 Evaluation of FSI simulation convergence with varying Young’s modulus
while Poisson’s ratio fixed to 0.35. The table categorizes the simulations
as either successfully converged (✓) or unable to converge, therefore, to be
completed (×). FSI = fluid-structure interaction; ν = Poisson’s ratio. . . . 21
5 Evaluation of FSI simulation convergence with varying Poisson’s ratio while
Young’s modulus fixed to 1 MPa. The table categorizes the simulations as
either successfully converged (✓) or unable to converge, therefore, to be
completed (×). FSI = fluid-structure interaction; E = Young’s modulus. . 21
6 Evaluation of FSI simulation stability with varying inlet velocity. The table
categorizes the simulations as either successfully completed (✓) or unable
to be completed (×). The scaled PS profile corresponds to the original
PS inlet velocity profile scaled by a factor of 0.1. FSI = fluid-structure
interaction; PS = patient-specific; vel. = velocity; E = Young’s modulus;
ν = Poisson’s ratio. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
7 Evaluation of FSI simulation stability with varying under-relaxation factor
values, reducing it from the default value of 1. The table classifies the simu-
lations as either successfully completed (✓) or unable to be completed (×).
FSI = fluid-structure interaction; E = Young’s modulus; ν = Poisson’s ratio. 22
A1 Values of the Ogden model parameters for modelling homogeneous artery
[57]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
1 Introduction
1.1 Background
Transcatheter aortic valve implantation (TAVI) is a minimally invasive heart proce-
dure that consists of replacing a thickened aortic valve that cannot fully open with
a prosthetic valve. This technique involves inserting a catheter into a blood vessel
through the femoral, the subclavian, or the carotid artery in the patient’s upper leg
or chest and passing it toward the aortic valve. Subsequently, the catheter is utilized
to steer and secure the replacement valve on top of the existing natural valve or to
perform a re-implantation of a prosthetic second valve. This is done by substitut-
ing the native aortic valves with the prosthetic ones that are enclosed within the
prosthesis, as seen in Figure 1.

Figure 1: Transcatheter aortic valve implantation (TAVI) procedure example. A new


aortic valve is placed over the top of the old one through a cardiac catheter. Retrieved
from [1].

TAVI has emerged as a promising alternative to conventional aortic valve replace-


ment for patients with severe, symptomatic aortic stenosis (AS). This is a common
heart pathology consisting of the aortic valve’s narrowing leading to an obstruc-
tion of the left ventricular outflow tract (LVOT). AS has been reported to be the
most common valvular heart disease with a prevalence of around 3% in the Spanish
population above 65 years, increasing the percentage up to 7.4% in older ages [2].
Traditionally, AS is treated via surgical aortic valve replacement (SAVR). TAVI was
considered an alternative intervention for high-risk patients who could not undergo
SAVR. However, given the low complication rates reported after undergoing the
TAVI procedure, it has been approved for all AS risk groups [3].

1
Nowadays, there is a wide range of commercially available transcatheter aortic
heart valves differing from each other. There are differences in design and, therefore,
in the mechanism of action, as seen in Figure 2. The mechanism of action of the
bioprosthetic leaflets can be intra-annular or supra-annular, and the valves can be
either ballon-expandable, self-expanding, or mechanically expanded. Regarding the
structure of the skirt, this is commonly encountered between commercial devices to
prevent paravalvular leaks. The frame size, height, and range designs differ based
on anatomical measurements and procedural feedback from interventional cardiol-
ogists [4]. Although any of these valves can be successfully implanted in patients,
the interventional and imaging cardiology team should be aware of the differences
between the distinct TAVI prosthesis models by understanding their respective ad-
vantages, disadvantages, and movement mechanism for different valve technologies
in adverse settings to achieve the best outcome depending on the clinical case.

Figure 2: Overview of families of bioprostheses for TAVI. The more transparent tran-
scatheter heart valves represent no longer available bioprostheses. Retrieved from [5].
TAVI = transcatheter aortic valve implantation.

Clinicians use a variety of computational tools to aid in their decision-making


during TAVI procedures. Some of these tools include imaging techniques such as
computed tomography (CT) and magnetic resonance imaging (MRI), which provide
detailed images of the patient’s aortic valve anatomy [6, 7]. These images are used
to evaluate the size of the native valve, measure the aortic annulus, and assess the
vascular access sites. Additionally, they use pre-planning software tools to view and
manipulate these images, allowing them to better understand the patient’s anatomy
and select the most appropriate TAVI device. Two commonly used software tools for
procedural planning for TAVI are 3 mensio (Pie Medical Imaging BV, The Nether-
lands) and HeartNavigator 3.0 (Philips, Amsterdam, The Netherlands) [8].

2
The 3 mensio software provides tools that allow a comprehensive assessment of
the implanted device landing zone and a detailed risk stratification regarding access
suitability. Although the 3 mensio tools are reliable, it is a semi-automatic tool
which, depending on the clinical context, can be time-consuming and lacks other
interesting pre-planning features. Therefore, HeartNavigator, which was recently re-
leased as a fully-automatic tool, emerged as a powerful tool for TAVI pre-planning in
clinical practice. HeartNavigator provides important planning measurements such
as annulus area and perimeter, coronary ostia distances, and optimal implantation
angle. However, HeartNavigator lacks features such as a tool for access site planning
and fluid hemodynamic information [8].

Both aforementioned tools are considered to be sophisticated, reliable and easy to


use with low intra- and inter-observer variability. Since HeartNavigator was released,
characterized by its automation and planning tools, it has become the preferred pre-
planning software used in daily clinical routines. Regardless of the limitations the
two software may present, these computational tools are crucial for minimizing the
risk of potential complications after the procedure [8]. However, with still ongoing
technical improvements in CT scans as well as software, procedural planning in
TAVI will be further facilitated to improve the outcomes.

Figure 3: Clinical images. HeartNavigator software being used for TAVI procedure plan-
ning. Retrieved from [9]. TAVI = transcatheter aortic valve implantation.

The steps of pre-procedural planning vary depending on the clinical case and the
hospital’s clinical protocol but they typically include:

1. Patient assessment, including a physical examination and review of medical


history to identify potential risks.
2. Imaging studies, such as CT or MRI, to evaluate the patient’s aortic valve
anatomy and determine the appropriate device size and access site [6, 7].
3. Device selection based on patient anatomy and measurements, with the aim
of minimizing complications.

3
4. Selection of the access site, taking into account patient anatomy, age, sex, and
potential risks.

5. Planning for potential complications, such as bleeding, thrombus formation,


atrioventricular (AV) block leading to permanent pacemaker implantation
(PPI), or stroke; by considering the patient’s medical history.

6. Coordination with other specialists, such as cardiologists and anesthesiologists.

7. Patient education prior to the procedure, including information on the proce-


dure, risks, and recovery period [10, 11].

As mentioned, in the pre-planning some of the potential complications that can


emerge from the TAVI intervention are addressed, but there are some limitations
that still can not be prevented with the currently available pre-planning tools. Some
of the post-TAVI related complications include paravalvular leaks (PVL), which
have been reported to occur in 7-40% of TAVI cases and can negatively impact both
short-term and long-term survival rates [12]; valvular calcifications, which have been
linked to adverse outcomes during and after TAVI [13]; delayed coronary obstruction
(DCO), which occurs when coronary arteries become blocked after TAVI [14]; the
need for PPI due to AV conduction disorders [15]; and device-related thrombus
(DRT) [16, 17].

1.2 State of the art


1.2.1 Device-realted thrombus
Device-related thrombus refers to the formation of blood clots (thrombi) specifically
in relation to implanted medical devices. It is a complication that can occur when
blood comes into contact with foreign surfaces, such as those found in medical de-
vices like stents, prosthetic heart valves, or artificial vascular grafts. The presence
of a medical device can disrupt the normal flow dynamics of blood, create areas of
stagnation or turbulence, and activate the coagulation cascade, leading to thrombus
formation.

Mechanisms underlying thromboembolic events after TAVI are multi-factorial.


Although the reasons behind DRT are not well understood, evidence suggests that
haemodynamic disturbances at sites of valve implantation play a leading role in
thrombus formation related to the placement of the prosthesis resulting in DRT.
However, it has been also proposed that it could be related to the design of the
implanted valve and to the implantation procedure techniques that could lead to
thrombus formation [18].

DRT is characterized by leaflet thrombosis post-TAVI, which has been found


to have an incidence of approximately 10-15% in patients who undergo CT an-
giography after TAVI [18]. Leaflet thrombosis can lead to reduced leaflet motion,
impede forward flow, elevate the transvalvular pressure gradient, increase the load
on the left ventricle; and possibly lead to heart failure, embolization, stroke and

4
death [19]. There may be case-specific factors that increase the risk of DRT such
as pre-existing thrombotic conditions, atrial fibrillation, coagulation disorders, or
inadequate anticoagulation therapy. Other factors such as the used valve type, im-
proper positioning of the prosthesis and inadequate expansion or under-sizing of the
valve can create regions of stasis or turbulence, predisposing to thrombus formation.
Procedural factors such as prolonged interventional duration, excessive manipula-
tion of the catheter or the valve, and inappropriate post-procedural anticoagulation
strategies may also contribute to thrombus formation.

The management of DRT after TAVI depends on the severity of the thrombosis,
the symptoms, and the underlying causes. Antithrombotic treatments post-TAVI
can range from a single or dual antiplatelet regimen to the use of oral anticoagu-
lants parallel to antiplatelet therapy [20]. In some cases, anticoagulation therapy
may be sufficient to prevent further thrombus formation and to allow the thrombus
to be dissolved [21]. However, research is needed to clarify the mechanism behind
the prothrombotic activity after TAVI and the optimal medical management, as the
anticoagulant treatment is still unclear [18].

Figure 4: Predictive model for thrombus formation after TAVI in the neo-sinus. Retrieved
from [22]. V = velocity; TAVI = transcatheter aortic valve implantation; TAV = tran-
scatheter aortic valves.

The imaging modalities commonly used to assess DRT after TAVI are trans-
esophageal echocardiography (TEE) and CT [8]. Both imaging techniques allow
clinicians to visualize important information about blood haemodynamics generated
at sites of valve implantation, leaflets, in the aorta, and surrounding vessels after the
intervention. Other imaging modalities such as MRI may also be used to evaluate
the presence of thrombi [23]. However, there are limitations to the mentioned in vivo
imaging after TAVI. TEE and CT are invasive techniques that require a high level of
skill and expertise to perform them. In addition, these imaging modalities may not
be able to detect small thrombi or other abnormalities in the blood flow patterns
near the leaflets, aorta and vessels. Furthermore, the low resolution of these imaging
modalities may not be sufficient to fully evaluate the complex blood flow patterns

5
that emerge in the anatomical structures surrounding the implanted device [23].
Therefore, these tools cannot provide the detailed measurements required to fully
investigate the mechanisms behind thrombus initiation and formation, particularly
in smaller anatomical regions such as the neo-sinus (refer to Figure 4).

1.2.2 Computational models


Given this lack of information from the imaging techniques regarding blood haemo-
dynamics post-TAVI, numerical simulations have emerged as alternative tools. These
are useful to predict and quantify the mechanical and fluid dynamics in the context
of having cardiovascular devices implanted, such as TAVI, and evaluate how they
interact with a patient’s specific heart anatomy.

Study Software TAVI prosthesis Aortic wall Inlet Outlet Cardiac cycles

Mao, W. [24] Abaqus Leaflets Rigid PS press. PS press. 2

Kandail, H. [25] FlowVision/Abaqus Frame, leaflets and skirt Rigid Lit. vels. Lit. press. 3

Luraghi, G. [26] ANSYS (LS-Dyna) Frame, leaflets and skirt Rigid PS press. PS press. 2

Basri, A. [27] ANSYS Frame, leaflets and skirt Linear elastic Mass flow Lit. Press. 3

Govindarajan, V. [28] 3D-FSI algorithm Leaflets Rigid Lit. vels. Lit. Press. 2

Borowski, F. [29] ANSYS Leaflets Rigid Av. vels. 0 Pa 1

Liu, X. [30] ANSYS Leaflets Rigid Av. vels. 0 Pa 1

Brown, J. A. [31] IBAMR Frame, leaflets and skirt Neo-Hookean Lit. wind. Lit. wind. 1

Oks, D. [32] Alya Leaflets Rigid Lit. flat vels. Conv. BCs 1

Ghosh, R. P. [33] FlowVision/Abaqus Leaflets Rigid Lit. press. Lit press. 3

Sadrabadi, M. S. [34] ANSYS Leaflets Linear elastic Lit. vels. 0 Pa 2

Vahidkhah, K. [35] ANSYS Frame and leaflets Rigid Lit. press. Lit press. 3

Fumagalli, I. [36] LifeV Frame and leaflets Rigid Lit. vels. Lit press. 1

Pil, N. [37] COMSOL Leaflets Rigid PS vels. 0 Pa 1

This work ANSYS Leaflets Linear elastic PS vels. PS press. 1

Table 1: Review of boundary conditions and different modelling choices for TAVI sim-
ulations reported in the literature. TAVI = transcatheter aortic valve implantation; FSI
= fluid-structure interaction; lit = literature; vels = velocities; press = pressures; PS =
patient-specific; wind = Windkessel; conv = convective; BCs = boundary conditions.

Computational fluid dynamics (CFD) simulations are run to solve the Navier-
Stokes equations for obtaining numerical solutions regarding pressure and velocity.
In this case, CFD studies are carried out to study the variation of blood flow pat-
terns in the aortic root because of the aortic valves implanted after TAVI. Several
clinical complications can be addressed with a CFD approach, such as PVL [38] and
DRT, by studying variations in the aortic flow patterns post-TAVI [39, 40]. However,

6
because of the strong interaction between the implanted valve and the surrounding
anatomical structures, fluid-structure interaction (FSI) analysis is considered to be
the best computational approach for accurate simulation of the valve load and the
surrounding flow field [26, 30]. In contrast with CFD, the FSI approach allows for
the representation of the dynamic movement of the anatomical structures during
the cardiac cycle.

There are different ways to address TAVI using FSI simulations. To minimize
computational time and resources, as well as to reduce model complexity, simplifica-
tions are made to both the patient-specific anatomical geometry and TAVI device.
The simplest model to study the prosthesis leaflet motion under the FSI approach
usually involves assuming the aorta as a rigid tube and modelling only the TAVI
leaflets as non-rigid material. Hence, the interaction between the blood flow and
the leaflets is studied [24, 29]. Degrees of complexity can be added by incorporat-
ing patient-specific anatomical structures such as the coronary arteries and native
valves, segmented from patient-specific CT images, while still only considering the
leaflets from the TAVI prosthesis [28]. Nevertheless, recently, more FSI studies have
been done that included all parts of the TAVI prosthesis: leaflets, skirt and metallic
frame, along with the patient-specific aorta model [31, 26], adding degrees of com-
plexity.

For the assessment of blood flow patterns and leaflet motion that may lead to
DRT, various parameters are analyzed including analysis and quantification of leaflet
rigidity and orifice area, which is an indicator of the valve’s degree of opening. The
effective orifice area (EOA) or the geometric orifice area (GOA) can be calculated
from flow measurements or the valve geometry [32]. Parameters such as wall shear
stress (WSS) [34], von Mises stress, shear stress [41, 42, 43], particle accumulation
(PA) quantification, vortex location identification (VLI), and transvalvular pres-
sures and velocities are also analyzed to study and evaluate blood flow dynamics
after implantation [25, 28]. These parameters provide numerical comparisons of
prosthesis performance and flow dynamics, enabling a thorough evaluation of the
risk of thrombus formation.

Figure 5: Cross-sectional view of detailed flow patterns for the TAVI Evolut R model.
The colour indicates the magnitude of the velocity through the aortic root, with cooler
colours indicating lower velocities and warmer colours indicating higher velocities. The
time increment between frames is 0.1s. Retrieved from [31]. U = velocity.

7
Although FSI is a powerful approach, published studies exploring the interaction
between the blood flow and the patient-specific anatomical structures including the
implanted device are limited, see Table 1. Most published investigations regard-
ing FSI analysis study leaflets dynamics assuming aortic walls as absolutely rigid.
Therefore, neglecting the elasticity of the aortic wall may have a considerable impact
on local haemodynamics. In a recent study [44], it was observed that assuming rigid
or elastic models resulted in a substantial difference, particularly immediately after
valve closure. The rigid wall model exhibited larger and longer oscillations in both
flow rate and valve movement compared to the compliant arterial wall model.

Given the wide array of modelling approaches reported in FSI studies related to
TAVI, there is a need to establish best practice guidelines for constructing robust
models and obtaining reliable simulations that accurately reflect relevant clinical
outcomes. Consequently, this study aims to propose a computational workflow that
incorporates an FSI approach to model the interaction between blood and patient-
specific anatomical structures following TAVI, while considering the non-rigid nature
of the aortic walls. The proposed workflow seeks to enhance the accuracy and
reliability of simulations in the context of TAVI.

1.3 Objectives
The main goal of this thesis is to present a computational methodology for mod-
elling patient-specific TAVI using an FSI approach assuming a non-rigid aortic wall
behaviour to investigate the factors leading to DRT after TAVI intervention. For
this purpose, a sensitivity analysis is presented to evaluate the relevance of the most
important modelling choices in TAVI FSI simulations. The proposed workflow con-
sists of the following steps: (1) acquiring patient-specific data; (2) developing the
model including patient-specific geometry and TAVI prosthesis; (3) setting up the
FSI simulation; (4) validating the results obtained; and (5) analyzing haemodynamic
thrombus indices to assess the risk of DRT after TAVI.

8
2 Methods
In this section, the steps followed to develop an FSI simulation in TAVI are presented
in Figure 6. The workflow involves: (1) segmentation of anatomical structures from
patient-specific CT images; (2) acquisition of Computer-Aided Design (CAD) files
for both fluid and solid domains, crucial for establishing mechanical properties prior
to meshing; (3) determination of material properties and patient-specific bound-
ary conditions for both fluid and solid domains; (4) evaluation of haemodynamic
thrombus indices.

Figure 6: Proposed modelling workflow established for TAVI simulation using FSI ap-
proach with a patient-specific geometry. TAVI = transcatheter aortic valve implantation;
FSI = fluid-structure interaction; CT = computed tomography; CAD = computer-aided
design.

2.1 Clinical data acquisition


Clinical data of one patient who was treated with a Navitor 25 mm (Abbott, Chicago,
USA) prosthesis, was provided by Hospital de la Santa Creu i Sant Pau (Barcelona,
Spain). This data comprised pre- and post-procedural cardiac CT images acquired at
systolic and diastolic phases with a Somatom Force (Siemens Healthineers, Erlangen,
Germany) scan. The CT images covered the outflow tracts from the left ventricle
to the mid-ascending aorta. In the case of the post-procedural CT images, the
allocated prosthesis was covered entirely.

2.2 Anatomical 3D model geometry


2.2.1 Patient-specific 3D model
The provided CT images in DICOM (Digital Imaging and Communications in
Medicine) format were imported into Mimics 25.0 software (Materialise, Leuven,
Belgium) to perform the three-dimensional (3D) patient-specific model reconstruc-
tion. A semi-automated segmentation algorithm based on thresholding and region
growing was used to segment the aortic root, coronary arteries and transcatheter
aortic valve (TAV) prosthesis using Mimics. Then, the 3D model obtained was ex-
ported into Stereo-Lithography (STL) format to be loaded in ANSYS SpaceClaim
22.1 (Ansys, Inc., Pennsylvania, USA) for the generation of a solid model.

9
2.2.2 TAVI prosthesis model
The design of the TAVI prosthesis aimed to encompass all its components, includ-
ing the leaflets, skirt, and metallic frame. SolidWorks software (Dassault Systèmes,
Suresnes, France) was utilized to design each of these parts individually. Subse-
quently, these parts were intended to be unified to obtain the final device in CAD
file format.

As the dimensions of the leaflets were not available, it was not feasible to create
them from scratch. Therefore, leaflets from a different TAVI prosthesis were utilized
and subsequently modified to align with the specifications of the new CAD genera-
tion. The design details for the metallic frame were obtained from studies [45, 35],
which provided comprehensive metallic frame design guidelines as seen in Figure 7;
for more details refer to Appendix A.1. However, due to time limitations and a lack
of literature information on its design, the skirt component was not included in the
design.

Figure 7: (a) TAVI prosthesis metallic frame cell retrieved from [46], (b) sketch frame
reconstruction in SolidWorks from provided cell design in (a), (c) obtained metallic frame
cell after reconstruction. TAVI = transcatheter aortic valve implantation.

2.2.3 Simplified 3D model


To demonstrate the proposed computational workflow, an initial proof-of-concept
was conducted using a simplified 3D model instead of the patient-specific 3D model.
The simplified 3D model was inspired by the patient-specific 3D model, incorporat-
ing assumptions from previous literature that employed simple models to investigate
prosthesis leaflet motion under an FSI approach [24, 29]. Hence, the simplified model
consisted of a tubular structure representing the aorta with TAVI leaflets attached,
refer to Figure 8. Regarding the dimensions and properties of the tubular structure,
they were defined using values of thickness, length, and mechanical properties as
described in literature [31].

10
Figure 8: Geometries used for the development of the simplified model, (a) TAVI leaflets
geometry recovered from [28], (b) representation of the aorta (tubular structure), (c) ob-
tained simplified model for simulation. TAVI = transcatheter aortic valve implantation.

In order to further simplify the TAVI prosthesis to minimize computational costs


and prosthesis complexity, only the leaflets were considered in the model, while the
metallic frame and skirt components were neglected. The TAVI leaflets, extracted
from a GitHub repository [28], were connected to the tubular structure using the
Boolean Union tool in Mimics. The final simplified model, shown in Figure 8,
was converted into STL format (.stl) before being exported to SpaceClaim for the
generation of a solid model.

2.3 Solid domain model


2.3.1 Conversion of STL file to CAD file format
In order to accurately define the physical properties of the solid model, the gen-
erated STL file was converted into a CAD file format. This conversion step was
important since the STL file format represents the 3D model’s surface as a mesh
composed of triangular faces, lacking the necessary information required for precise
modelling and analysis, specifically in the context of defining the physical and me-
chanical properties of the solid domain model.

To achieve the proposed STL to CAD transformation, first, Meshmixer (Au-


todesk, Inc., California, USA) was used to remesh the model, add thickness to the
leaflets, and smooth the model’s surface; ensuring a highly accurate conversion from
the reference STL file to CAD. In SpaceClaim, the automatic Auto Skin tool was
employed to convert the adjusted STL file to CAD format. Manual corrections were
iteratively performed on both the STL file in Meshmixer and the CAD file in Space-
Claim to eliminate any mesh errors and create an STL replica that closely resembled
the original CAD file (see Figure 9).

11
Figure 9: Conversion from STL to CAD. (a) STL file of reference to which several manual
corrections were applied in Meshmixer. (b) Converted STL to CAD file using ANSYS
SpaceClaim 22.1 tools of correction and conversion. STL = stereo-lithography; CAD =
computer-aided design.

Post-processing tools were employed to perform modifications to refine the CAD


model in SpaceClaim. The aorta tube representation was adjusted to a diameter,
length, and thickness of 26 mm, 76.8 mm and 2 mm, respectively [31]. Similarly, a
constant thickness was defined in the leaflets until they reached a uniform thickness
of 0.4 mm [26, 31]. Furthermore, the geometry was divided into distinct parts to
define the regions that would interact with the fluid domain. In order to prevent
the geometry from displacing, movement restrictions were imposed. Specifically, the
inlet and outlet edges were defined as fixed supports in all three dimensions.

2.4 Fluid domain model


The fluid domain represented blood. Since blood flows inside the aorta and interacts
with the leaflets and the aortic walls, the solid model geometry was used as a tem-
plate to generate the fluid model. Therefore, the fluid’s geometry was enclosed by
the solid. For this process, the used software was SpaceClaim. The Volume Extract
tool was employed to generate the fluid domain based on the enclosing solid volume.
The resulting geometry was the representation of the fluid (see Figure 10b) which
fitted perfectly with the solid model (see Figure 10c).

Figure 10: Obtained CAD files. (a) Solid domain geometry. (b) Fluid domain geometry.
(c) Both fluid and solid geometries where the fluid geometry is enclosed within the solid
geometry. CAD = computer-aided design.

12
2.5 Boundary conditions
Boundary conditions were imposed by defining the inlet of the system as the lower
part of the aorta, representing the LVOT, and the outlet as the upper part of the
aorta, representing the ascending aorta. To differentiate between the top and bot-
tom boundaries, the orientation of the leaflets was considered as a reference point.

The blood flow velocity curve at the LVOT was obtained using Doppler echocar-
diography, as described in Appendix A.2. This velocity curve was then utilized as
the input condition for the model. For the outlet boundary condition, a constant
pressure of 100 mmHg was set at the ascending aorta. This value was chosen based
on the literature [47], and approved by clinicians from Hospital Sant Pau i Santa
Creu. This assumption ensured consistency with clinical practices and provided a
reliable boundary condition for the simulation.

2.6 FSI simulation setup


2.6.1 Solid structure analysis
The structural analysis was performed using ANSYS Mechanical Solver 22.1 (Ansys,
Inc., Pennsylvania, USA). Transient dynamic analysis in finite element analysis was
employed to investigate the structural behaviour of the solid domains under applied
loads.

The aorta was assumed to exhibit a linear elastic behaviour, with Young’s mod-
ulus (E) of 1 MPa and a Poisson’s ratio (ν) of 0.35 [26, 32, 27]. Regarding the
mechanical properties of the TAVI prosthesis leaflets, they vary depending on the
type of material used for their manufacturing process. They are generally made of
porcine pericardium or thin glutaraldehyde-treated bovine pericardium [24]. Given
that in this study the patient was treated with the Navitor prosthesis, its leaflets
were assumed to be made from bovine pericardial tissue, based on [46].

According to [24], bovine pericardium tissue used for the manufacturing of the
TAVI prosthesis leaflets can be assumed to be an incompressible, anisotropic, non-
linear, hyperelastic material; thus its strain energy function, W , can be expressed
by a fibre-reinforced hyperelastic material model (MHGO) based on the work of
Holzapfel, Gasser and Ogden [48, 49]; described in equation 1. In this model, leaflets
were assumed to be composed of a matrix material with two families of embedded
fibres, each consisting of a preferred direction [24].
2
C01 (I¯1 −3) κ1 X κ2 (I¯4i −1)2 1
W = C10 {e − 1} + {e − 1} + (J − 1)2 , (1)
2κ2 i=1 D

where C10 , C01 , κ1 , κ2 and D are material constants, I¯1 and I¯4i are the deviatoric
strain invariants. Both variables C10 and C01 are used to describe the matrix ma-
terial. Variable D refers to the material constant to impose incompressibility, and
J is the determinant of the deformation gradient. Finally, κ1 is a positive constant

13
with the dimension of stress to describe the fibre material and κ2 is a dimensionless
parameter. In Table 2 the values of material properties are presented.

C10 (kPa) C01 κ1 (kPa) κ2 Θ(◦ ) D (kPa−1 )


30.03 3.47 74.5 63.19 43.11 1.00e−5

Table 2: Glutaraldehyde-treated bovine pericardium material properties. Retrieved from


[24].

As an initial approach to the FSI simulation, the aortic walls and leaflets were
assumed to exhibit linear elastic behaviour. They were assigned values of: E = 1
MPa and ν = 0.35 [26, 32, 27]. This simplified assumption allowed for the explo-
ration of the FSI behaviour while considering a linear elastic response for the solid
components.

2.6.2 Fluid dynamics analysis


Fluid dynamics analysis was performed in ANSYS Fluent Solver 22.1 (Ansys, Inc.,
Pennsylvania, USA). Simulations were made in transient formulation under a lam-
inar model. The laminar model was used since the shear rate had values above 10
s-1 , and thus the flow could be considered as Newtonian [50].

Blood was modelled as a viscous, isothermal, incompressible and Newtonian fluid


using the Navier-Stokes equations. The momentum conservation and continuity
equations are defined by equations (2) and (3), respectively. Blood density and
viscosity were assumed to be ρ = 1060 kg/m3 and µ = 0.004 Pa·s, respectively [25].

∂v
ρ + ρ(v · ∇)v = ρg − ∇p + µ∇2 v (2)
∂t
∂ρ
+ ∇ · (ρv) = 0, (3)
∂t
where ρ is the density of the fluid, v is the velocity vector field, t is time, g is
the gravitational acceleration vector, p is the pressure, µ is the dynamic viscosity,
∇ is the gradient operator, and ∇2 is the Laplacian operator.

2.6.3 Fluid-structure interaction coupling


There are two different approaches to solving FSI problems, the monolithic ap-
proach and the partitioned approach. In the monolithic approach, the solid and
fluid domains are formulated as one combined domain, meaning that the interaction
between the fluid and solid at the interface is treated synchronously. This leads to
the conservation of properties which increases the stability of the solution. This
approach is considered to be more robust than the partitioned one. However, it is
computationally expensive and cannot take advantage of software modularity as the
partitioned method does [51].

14
As for the partitioned approach, solid and fluid domains are solved separately.
This means that the flow does not change while the structural solution is being
calculated. The equations governing the flow and the displacement of the solid are
solved alternatively in time with two different solvers [51]. Therefore, the interme-
diate fluid solution is prescribed as a boundary condition for the structure and vice
versa; the iteration continues until the convergence criteria are satisfied. At the
boundary between fluid and solid, the exchange of information occurs according to
the type of coupling analysis applied [51, 52]. The coupling can be of two types:
one-way or two-way, as shown in Figure 11.

Figure 11: (a) One-way and (b) two-way coupled FSI approaches. Retrieved from [52].
FSI = fluid-structure interaction.

In the one-way coupling, the motion of the fluid flow influences the solid struc-
ture but the reaction of the solid upon the fluid is negligible. This allows the fluid
and solid analysis to be solved independently with unidirectional data transfer. The
other way around is also possible. However, in the two-way coupling, the motion
of the fluid influences the solid structure and simultaneously the flow of fluid is in-
fluenced by the reaction of the solid structure. Therefore, fluid and solid domains
are solved simultaneously with bidirectional data transfer. Force is exported from
the fluid to the solid domain, and deformation is transferred from the solid to the
fluid domain. This is done to update the mesh of the fluid domain every coupling
iteration until both solutions converge [51, 52].

In this study, the partitioned approach was employed and the interaction between
fluid and solid was coupled in a two-way mode in ANSYS Workbench 22.1 (Ansys,
Inc., Pennsylvania, USA). The two-way FSI analysis was performed by connecting
the coupling participants to a component system called System Coupling, which fa-
cilitates multidisciplinary simulations between coupling participants. A participant
system is a system which either feeds or receives data in a coupled analysis. In this
case, Transient Structural (participant 1) and Fluid Flow (participant 2) acted as
coupling participants, see Figure 12.

15
Figure 12: Two-way FSI coupling between Transient Structural and Fluid Flow in ANSYS
Workbench 22.1. FSI = fluid-structure interaction.

During the coupling process between the solid and fluid domains, data is ex-
changed between these entities. A data transfer involves the movement of a specific
variable type in one direction between two participants [53]. In this particular sce-
nario, two data transfers were defined. In the first transfer, the Transient Structural
component acted as the source, while the Fluid Flow component served as the tar-
get. The variable transferred between them was the incremental displacement. In
the second data transfer, the Fluid Flow component functioned as the source, while
the Transient Structural component acted as the target. The variable exchanged
between them was force.

Based on the findings in [53] regarding solution stability, certain multiphysics


scenarios, particularly those involving fluid-solid interaction with pressure/force and
displacement transfers, have a tendency to exhibit instability. These cases commonly
involve thin structures, as well as structural materials with a low Young’s modulus,
or liquids. In the current study, the variables being transferred were forces and
displacements, thin structures such as aortic walls and leaflets were considered, and
aortic walls were defined with a low Young’s modulus; all of which possess the prop-
erties known to contribute to potential instabilities.

There are techniques in System Coupling to help overcome solution instabilities


which include under-relaxation, ramping, and artificial compressibility. In the cur-
rent study, only the under-relaxation factor was tuned, the other two were set as
default. The under-relaxation factor is a constraint on the change of target data
from one coupling iteration to the consecutive and it is restricted to be between
0 and 1. Setting the under-relaxation factor below 1 could help stabilize coupled
response, however, it may lead to a deceleration of the simulation [53].

The convergence of data transfers is evaluated by comparing each coupling it-


eration with the previous one. When the normalized value between two successive
iterations falls below the convergence target, the transferred data is considered con-
verged. A coupling step is considered converged only when both the data transfers
and the individual solvers have achieved convergence. If a coupling step is not yet

16
converged, a new coupling iteration is initiated, unless the maximum number of
coupling iterations has been reached. In this study, various iterations ranging from
80 to 400 were tested to determine the optimal number of iterations. The selection
of the optimal number depended on the specified material properties and boundary
conditions.

Finally, for both the solid and fluid domains the time step was set to 0.001s,
simulation end time was set to 1 s and a second-order implicit formulation was
employed in the fluid domain; residuals for continuity equations were set as 0.001
for convergence criteria. The simulations lasted 1 beat.

2.7 Sensitivity analysis


This section comprises four sensitivity analyses conducted to assess the fluid-solid
interaction considering variations in different parameters. These four parameters
are: the mesh element size, mechanical properties assigned to the solid domain,
inlet velocity, and the under-relaxation factor value.

2.7.1 Mesh convergence study


Meshing was completed for both the solid and fluid components in ANSYS Meshing
22.1 (Ansys, Inc., Pennsylvania, USA). To investigate the impact of both the ele-
ment size and the mesh element number on the solver solution, a mesh convergence
study was conducted, by changing the range element size and evaluating the total
mesh elements for the solid and the fluid domains geometries separately. The same
case was run using the same boundary conditions with the only variation being the
element size, therefore, the number of elements present in the solid and fluid meshes.

The selected meshing technique utilized tetrahedrons as the primary method to


generate a complete tetrahedral mesh. The meshing algorithm employed was the
Patch Independent method, known for its effectiveness in handling complex CAD
geometries with significant irregularities, such as the leaflets. The same meshing
criteria were consistently applied to both the solid and fluid domains.

To assess the mesh quality, seven tests were conducted, varying the range of
element sizes. The total number of mesh elements for the solid and fluid domains
was evaluated separately during this analysis.

2.7.2 Mechanical properties


To examine the influence of key parameters such as Young’s modulus and Poisson’s
ratio on the rigidity of the aortic wall and leaflets, a variation in solid domain prop-
erties was simulated, consequently affecting the interaction with the fluid domain.
As previously mentioned, the aortic wall and leaflets were characterized as linearly
elastic materials, with a Young’s modulus of 1 MPa and a Poisson’s ratio of 0.35. A
progressive sequence of experiments was performed to explore the behaviour of the

17
system under the changes in Young’s modulus and Poisson’s ratio, individually.

The Young’s modulus values were incrementally adjusted according to the set:
E = {2, 3, 5, 10, 20, 30} MPa, while the Poisson’s ratio remained constant at 0.35.
To evaluate the impact of the Poisson’s ratio on the model’s performance, Young’s
modulus was fixed at 1 MPa, while the Poisson’s ratio values were adjusted according
to the set: ν = {0.3, 0.4, 0.45, 0.49}. The aforementioned tests were conducted under
identical boundary conditions, with a focus on examining the impact of different
levels of rigidity within the solid structure.

2.7.3 Velocity analysis


To investigate the influence of inlet velocity magnitude on the deformation of the
aortic wall and leaflets, a set of velocities were evaluated starting from low con-
stant inlet velocities of 0.001 m/s to the patient-specific inlet velocity profile, which
reached its peak velocity at 1.25 m/s. For all tests conducted at different velocities,
both the aortic wall and leaflets were defined as linear elastic materials with Young’s
modulus of 1 MPa and a Poisson’s ratio of 0.35.

2.7.4 Under-relaxation factor analysis


During the fluid-solid system coupling, simulations led to divergence when using the
default under-relaxation factor, which is 1. As the under-relaxation factor deter-
mines the rate at which the coupling between different physical domains is updated,
a high value can lead to instability in the solution process. This instability mani-
fests as abrupt changes or oscillations in the coupling system progress, resulting in
simulation failure.

To address these errors and ensure simulation convergence, it was necessary to


reduce the under-relaxation factor. By lowering the factor, the exchange of informa-
tion between the fluid and solid domains was slowed down, allowing for smoother
convergence and effectively preventing simulation instabilities. For that, a series of
test simulations were performed with varying under-relaxation factors to determine
the optimal under-relaxation factor value. In these tests, the inlet velocity was de-
fined as the scaled PS velocity profile and both the aortic walls and leaflets as linear
elastic materials.

2.8 Mechanical and haemodynamic indices


To assess the risk of DRT following TAVI, several mechanical and haemodynamic
indices were computed. The visualization and post-processing of these indices were
carried out using ANSYS Mechanical and Paraview 5.4.16 (Kitware, Inc., New York,
USA).

Since DRT after TAVI is associated with low blood velocities due to blood sta-
sis in the neo-sinuses [54], velocity was computed and analyzed in these regions.
Re-circulation velocity patterns could also lead to endothelial lesions and thrombus

18
formation due to the activation of platelets [41], hence, they were also analyzed.
FSI-derived flow streamlines were visualized to assess these conditions, focusing on
regions near the neo-sinuses with re-circulations and low velocities. A legend thresh-
old of 0.2 m/s was set, and the flow direction was represented using arrows. As for
mechanical indices, shear stress was computed to study the association between low
shear stresses and platelet aggregation, known to contribute to thrombus formation
[42]. Principal maximum stress and von Mises stress were calculated to study stress
distribution resulting from pressure loading on the leaflets. Regions with peak or
locally higher stresses were identified as areas prone to initiating tissue prosthesis
degeneration, affecting blood flow dynamics and potentially leading to DRT [43].
Displacement distributions of the leaflets, influenced by fluid forces, were also evalu-
ated. These FSI-derived mechanical indices were evaluated at each simulation time
step.

19
3 Results
3.1 Sensitivity analysis
This section presents the results of the four sensitivity analyses conducted to evalu-
ate the fluid-solid interaction considering variations in different parameters. These
sensitivity analyses aimed to identify the parameters that contribute to the conver-
gence of the FSI simulations. By examining the impact of some parameters, valuable
insights were gained regarding their influence on the simulation results and their role
in achieving convergence of the simulations.

3.1.1 Mesh convergence study


In Table 3, meshing analyses are reported. Seven element size ranges were explored,
ranging from 0.2 - 1 mm up to 3 - 5 mm, to determine the optimal meshing con-
figuration. The generation of meshes was carried out for each element size range,
considering both solid and fluid domains. However, certain ranges presented chal-
lenges, as meshes were unable to be generated accurately because of the geometry
irregularity and complexity. Specifically, for the element size ranges of 0.2 - 1 mm,
1 - 4 mm, 2 - 4 mm, and 3 - 5 mm, either the fluid mesh, solid mesh, or both
meshes could not be generated preserving the CAD geometry. Surface discontinu-
ities appeared in the leaflets within these element size ranges, primarily due to the
thinness of the leaflet surface and the likelihood of holes forming between the upper
and lower leaflet surfaces (see Figure A4). Nonetheless, meshes that preserved the
CAD geometry were successfully generated for element size ranges of 0.5 - 2 mm, 1
- 2 mm, and 1 - 3 mm.

Test Range element size (mm) Solid mesh elements Fluid mesh elements

1 0.2 - 1 652k 4.25M

2 0.5 - 2 208k 404k

3 1-2 163k 192k

4 1-3 160k 171k

5 1-4 158k 151k

6 2-4 46k 54k

7 3-5 22k 30k

Table 3: Review of mesh converge analysis performed over the solid and fluid domains in
ANSYS Meshing 22.1.

3.1.2 Mechanical properties


In Table 4, Young’s modulus values tested are reported, while maintaining the Pois-
son’s ratio at 0.35. The corresponding simulation convergence was reported to in-
dicate whether the simulations were executed successfully or encountered computa-

20
tional issues, resulting in their inability to be completed. It is worth noting that,
except for the 30 MPa value, all other tested values resulted in simulation failure.
A similar analysis was conducted and reported in Table 5, where different Poisson’s
ratio values were tested while Young’s modulus was fixed to 1 MPa. In this analysis,
none of the tested Poisson’s ratio values led to simulation convergence.

Young’s modulus (MPa) - ν = 0.35 Simulation convergence

1 ×

2 ×

3 ×

5 ×

10 ×

20 ×

30 ✓

Table 4: Evaluation of FSI simulation convergence with varying Young’s modulus while
Poisson’s ratio fixed to 0.35. The table categorizes the simulations as either successfully
converged (✓) or unable to converge, therefore, to be completed (×). FSI = fluid-structure
interaction; ν = Poisson’s ratio.

Poisson’s ratio - E = 1MPa Simulation convergence

0.1 ×

0.2 ×

0.3 ×

0.35 ×

0.4 ×

0.45 ×

0.49 ×

Table 5: Evaluation of FSI simulation convergence with varying Poisson’s ratio while
Young’s modulus fixed to 1 MPa. The table categorizes the simulations as either suc-
cessfully converged (✓) or unable to converge, therefore, to be completed (×). FSI =
fluid-structure interaction; E = Young’s modulus.

21
3.1.3 Velocity analysis
In Table 6, an overview of the evaluated inlet velocities and their impact on sim-
ulation convergence is presented. The table indicates whether the simulations suc-
cessfully converged or encountered computational errors. The results of the tested
velocities showed that when setting constant velocity magnitudes and scaled PS ve-
locity profile, the simulations converged successfully. However, when the original PS
velocity profile was tested, simulations failed.

Inlet velocity (m/s) - E = 1 MPa, ν = 0.35 Simulation convergence

0.001 (constant) ✓

0.01 (constant) ✓

0.1 (constant) ✓

Scaled PS profile (vel. profile; 0.1 x PS profile) ✓

PS profile (vel. profile) ×

Table 6: Evaluation of FSI simulation stability with varying inlet velocity. The table
categorizes the simulations as either successfully completed (✓) or unable to be completed
(×). The scaled PS profile corresponds to the original PS inlet velocity profile scaled by a
factor of 0.1. FSI = fluid-structure interaction; PS = patient-specific; vel. = velocity; E
= Young’s modulus; ν = Poisson’s ratio.

3.1.4 Under-relaxation factor analysis


Table 7, reports the tested under-relaxation factor values along with their impact
on the simulation, determining whether it led to failure or convergence. Five under-
relaxation factor values were tested, gradually decreasing from 0.75 to 0.35 with an
interval of 0.1 between consecutive values. The simulation convergence threshold
was reached at a value of 0.45.

Under-relaxation factor - E = 1 MPa, ν = 0.35 Simulation convergence

0.75 ×

0.65 ×

0.55 ×

0.45 ✓

0.35 ✓

Table 7: Evaluation of FSI simulation stability with varying under-relaxation factor val-
ues, reducing it from the default value of 1. The table classifies the simulations as either
successfully completed (✓) or unable to be completed (×). FSI = fluid-structure interac-
tion; E = Young’s modulus; ν = Poisson’s ratio.

22
3.2 FSI simulation results
Based on the sensitivity analysis, parameters that facilitate a stable outcome of the
FSI simulation were identified. Hence, it was decided to consider a physical scenario
with the studied parameters to evaluate the feasibility of the obtained FSI simulation
outcome. The FSI simulation setup included the following specific parameters:

1. Meshing element size ranging from 1 to 3 mm.

2. Inlet: PS velocity profile scaled by a factor of 0.1.

3. Outlet: constant pressure of 100 Pa.

4. Aortic walls and leaflets described by a linear elastic model with a E = 1 MPa
and a ν = 0.35.

5. Under-relaxation factor set to 0.45.

In the same line, based on the sensitivity analysis conducted, an additional sim-
ulation setup was established. This FSI simulation was set to test the behavior of
a second-order Ogden model in the solid domain, as opposed to the linear elastic
model. Detailed information regarding the FSI simulation setup and the obtained
results can be found in Appendix A.4.

3.2.1 Mechanical and haemodynamic indices


This section presents the obtained results for velocity, displacement distributions,
maximum principal stress, von Mises stress, and shear stress. Figure 13 shows a
comparison of the fluid flow streamlines extracted from the fluid domain at four
time steps: 0.25, 0.5, 0.625, and 0.75s. Flow re-circulation patterns were observed
at low velocities ranging from 0.03 to 0.1 m/s during leaflet peak opening (t = 0.5s),
as well as during leaflet closing phases (t = 0.625s and t = 0.75s).

In terms of the computed mechanical indices, Figure 14 presents the displace-


ment distributions (mm) observed in the leaflets. Significant displacements were
observed on the outer rim of the leaflets, particularly in areas directly influenced
by the flow. The maximum displacement was recorded during the leaflet opening
phase at t = 0.5s, reaching 0.5158 mm. Regarding mechanical stresses, Figure 15
shows the maximum principal stress (MPa) registered in the leaflets. Stresses rang-
ing from 0.00063 to 0.0038 MPa were observed on the leaflets upper surface during
the opening and closing phases, with accentuated stress levels during peak leaflet
opening.

23
Figure 13: FSI-derived blood flow velocity streamlines at four instances. The first row (a-
d) shows the top view of the leaflets opening and closing kinematics at the corresponding
instances. The second row shows the cross-sectional view of detailed flow patterns. The
colour indicates the magnitude of the velocity through the aortic root at the centre plane
of the valve, with cooler colours indicating lower velocities and warmer colours indicating
higher velocities. Re-circulations at low velocities are marked down with a black circle.
FSI = fluid-structure interaction; t = time.

Figure 14: Leaflet kinematics and FSI-derived total displacement (mm) at six instances
during leaflet opening and closing. FSI = fluid-structure interaction.

24
Figure 15: Leaflet kinematics and FSI-derived maximum principal stress (MPa) at six
instances during leaflet opening and closing. FSI = fluid-structure interaction.

25
4 Discussion
The main goal of the present study was to propose a computational workflow for
assessing patient-specific DRT risk after TAVI using an FSI approach. Despite sev-
eral TAVI-based FSI models available in the literature, there is a lack of a consensus
on the most appropriate set of mechanical properties for the aortic walls modelling
and used boundary conditions for simulations setup, refer to Table 1. Therefore, in
this work, a computational workflow and a sensitivity analysis have been presented
to demonstrate the relevance of the most important modelling choices in TAVI FSI
simulations for the prediction of DRT after implantation.

In this regard, four sensitivity analyses were performed to assess fluid-solid in-
teraction under parameter variations. Starting from the mesh convergence study,
results obtained indicated that three ranges of element sizes generated meshes that
accurately preserved the CAD geometry. The element size ranging from 1 - 3 mm
was chosen for meshing yielding the most reliable results in terms of computational
cost and time. Nevertheless, further analysis is necessary to determine whether the
meshing criteria can be extrapolated to a patient-specific 3D model.

Tests performed varying mechanical properties, refer to Tables 4 and 5, showed


that only the simulation conducted with the largest Young’s modulus value, 30
MPa, was completed. Simulations with Young’s modulus values of 1, 2, 3, 5, 10,
and 20 MPa were unable to reach completion due to stability issues associated with
the solid domain. These results indicate that an increase in Young’s modulus led
to simulation convergence. Regarding the evaluated values of Poisson’s ratio, while
maintaining Young’s modulus to 1 MPa, it was demonstrated that none of the tested
Poisson’s ratio values resulted in simulation convergence. Hence, changing Poisson’s
ratio without modifying Young’s modulus did not improve the simulation conver-
gence.

As for velocity analysis performed, refer to Table 6, it was shown that three con-
stant values of inlet velocity, 0.001, 0.01 and 0.1 m/s resulted in successful simulation
convergence. The simulation also converged when using the scaled PS velocity pro-
file. However, computational errors arose when testing with the original PS velocity
profile. These errors were associated with the solid domain, leading to simulation
failure during the initial coupling iterations.

Finally, the threshold under-relaxation factor value that ensured simulation con-
vergence without encountering instability was 0.45, see Table 7. While lower values
such as 0.35 could also be considered, it should be noted that they may lead to a
higher accumulation of errors. Therefore, the chosen under-relaxation factor value
was 0.45. This value led to a slower convergence of the simulation and enhanced its
stability in this particular case. Nevertheless, it should be emphasized that reduc-
ing the under-relaxation factor can lead to a loss of transferred data or a decrease
in the accuracy of the solution. This fact is produced due to the dampened in-
formation exchange between the fluid and solid domains, potentially impacting the

26
fidelity of the simulation results. Therefore, it is important to carefully consider the
trade-off between stability and accuracy when adjusting the under-relaxation factor.

As a consequence of the results obtained from the sensitivity analysis, some pa-
rameters that contribute to the simulation convergence were identified. Hence, a
physical scenario considering the analysed parameters was set up to study the FSI-
driven fluid flow dynamics within the solid structure interaction. The coupling study
was done by computing mechanical and haemodynamic indices in order to assess
DRT after TAVI. An additional FSI simulation was performed under a second-order
Ogden material defined for the solid domain, refer to Appendix A.4. This analysis
resulted in simulation failure during advanced coupling iterations. However, the
registered results before simulation failure (refer to Figure A5) revealed an increase
in the elasticity of the fluid domain aortic walls, phenomena not observed with the
linear elastic behaviour (as shown in Figure 13). Leaflets themselves did not undergo
any changes from their initial configuration in the Ogden model (as depicted in the
top row of Figure A5), whereas in the case of linear elastic behaviour, the leaflets
deformed at each instance (as illustrated in the top row of Figure 13).

There is an important condition to consider regarding the results. These are


based on a proof-of-concept FSI simulation setup after sensitivity analyses and cali-
brations were studied. Therefore, these results do not pretend to determine whether
the case study would develop DRT or not, but to evaluate whether the obtained
results are consistent with the simulation setup and, qualitatively evaluate if, at its
scale, fluid velocity and mechanical stress distribution are consistent with previous
studies.

Thrombus formation is more likely in low-flow regions with re-circulating flows


in the neo-sinuses, as mentioned previously. Figure 13 shows re-circulations at low
velocities (0.03 - 0.1 m/s) during peak leaflet opening (t = 0.5s) and closing (t =
0.625s and t = 0.75s). Figure 14 indicates significant leaflet displacements, reaching
its peak in the periphery where it interacts directly with the flow. The largest dis-
placement of 0.5158 mm occurs during leaflet opening at t = 0.5s, coinciding with
the highest inlet velocity of 0.125 m/s. As for the mechanical stresses, high-stress
regions registered on the leaflets reflect those areas that could be prone to initiating
degeneration and affecting flow dynamics [43, 55]. Mechanical stresses, shown in
Figures 15 and A6, distributed across the whole leaflet, reveal high-stress regions
on the leaflets that could initiate degeneration and impact flow dynamics [43, 55].
Peak stresses occur at the leaflet periphery, particularly at the upper edge and at
the attachment region to the aorta, consistent with previous studies [43, 31]. In
short, low-velocity re-circulations in the neo-sinuses align with low-stress regions,
while peak stresses at the leaflet periphery coincide with the zone of maximum dis-
placement.

Despite the obtained results can not lead to a quantitative thrombotic evaluation,
they demonstrate the power of using an FSI approach instead of other approaches
such as CFD. In addition to the velocity and WSS analyses typically obtained from

27
CFD, an FSI approach enables the study of valuable mechanical indices that allow
for the evaluation of clinically relevant performance metrics. These metrics surpass
the capabilities of conventional imaging techniques and include parameters that are
challenging to measure directly in vivo or in vitro, such as mechanical stress distri-
butions [31]. The evaluation of the results primarily focused on the leaflets due to
their direct interaction with blood flow and resulting motion. Therefore, it has been
demonstrated that the motion of the leaflets is possible without requiring to impose
pre-determined displacements on them, through the analysis of this leaflet-blood in-
teraction. In contrast, in a CFD study, the leaflet’s movement only could be reached
by defining a displacement function between leaflets opened and closed positions (ob-
tained from dynamic CT images) to generate movement of mesh boundaries during
the simulations. This process is manual, time-consuming, and reliant on factors such
as the dynamic CT resolution, segmentation expertise and the movement function
defined to reproduce the physiological movement. However, in the FSI approach,
the movement of the leaflets is governed by other factors such as the thickness of the
leaflets, the transfer of fluid forces to the solid structures, the mechanical properties
defined for the leaflets and the aorta, the prescribed boundary conditions, and the
chosen under-relaxation factor. The dependence of leaflet motion on many external
parameters introduces complexity to the problem, as model calibration is required to
determine the most suitable simulation setup to ensure physiological leaflet motion.
This complexity may explain the inconsistent findings in the literature regarding the
choice of boundary conditions and mechanical properties as depicted in Table 1.

4.1 Limitations
It has to be noted that the current study presents some limitations. Firstly, the
simplified model geometry employed for the simulations does not consider either
the patient-specific anatomy or the inclusion of the TAVI prosthesis, except for the
leaflets. This limitation may impact assessing how the whole TAVI prosthesis inter-
acts with the patient-specific anatomical structures of the patient. For this reason,
a more robust 3D model should be considered.

During the generation of the TAVI prosthesis CAD, several limitations were en-
countered due to the lack of available literature on the design of certain prosthesis
components. As a result, the skirt could not be generated, and the leaflets had to be
adapted from another TAVI CAD prosthesis. Additionally, the task of CAD design
required specific skills, resulting in a time-intensive process. Consequently, due to
time constraints and simplifications applied to the 3D model used for simulations,
the TAVI prosthesis CAD remained incomplete. Nevertheless, the generated metal-
lic frames and leaflets can be utilized for future work.

When considering the STL to CAD conversion procedure, it is important to


acknowledge that it was not a straightforward process and required manual correc-
tions, resulting in the loss of the original dimensions of the leaflets. Furthermore, the
tools available in SpaceClaim were limited, which further prolonged the process and
compromised the quality of the obtained CAD geometry such as preserving leaflet

28
features. Consequently, this could pose challenges in the meshing, particularly when
employing small element sizes, as observed in the meshing sensitivity analysis.

As for the mechanical properties, the initial purpose was to model the aortic
walls with a linear elastic material and leaflets with a fibre-reinforced hyperelastic
material. However, due to the complexity and time constraints, both the aortic walls
and leaflets were assumed to exhibit linear elastic behaviour during the study. This
assumption might have affected the physical resolution of the problem by leading
to simulation failures, as seen with the mechanical properties sensitivity analysis for
linear elastic and also for the second-order Ogden material test. For that, a potential
consideration for future work is the implementation of a multi-material approach.

Regarding the under-relaxation factor value of 0.45 could be increased to min-


imize the loss of transferred data and increase the solution accuracy. Therefore,
future work could involve exploring other parameters such as ramping and artificial
compressibility to help overcome solution instabilities without necessarily implying
the loss of transferred data.

Despite all the experiments carried out, none of the simulations performed with
PS boundary conditions and mechanical properties defined initially arrived at the
established total simulation time without failing. Problems could originate from
different sources such as the material parameters, boundary conditions, or even
meshing. However, due to the non-linear nature of the problem, FSI simulations
require many model calibrations. Hence, another possible source of error could be
that the numerical method used is not coping with the high non-linearity of the
problem.

A significant limitation of this study is the extensive computational resources


and time required for simulations. These factors constrained the number of parame-
ters that could be explored in the sensitivity analysis, limiting the ability to identify
an optimal set of parameters for simulating physiological velocities, patient-specific
models, and TAVI prostheses. The time-consuming nature of the simulations, com-
bined with resource constraints, posed challenges in achieving a comprehensive in-
vestigation of these complex scenarios. Consequently, due to time constraints, the
simulation results for the complete PS velocity profile and linear elastic behaviour
(E = 30 MPa, ν = 0.35) are still pending.

4.2 Future work


The work presented is a demonstration of the potential power of FSI simulations in
assessing performance and risks following TAVI. As described in the objectives, our
objective was to propose a computational workflow that utilizes the FSI approach
to model TAVI and investigate factors contributing to DRT after prosthesis implan-
tation. While we have introduced the computational workflow, there is room for
improvement to enhance its efficiency.

29
Firstly, incorporating patient-specific 3D models and the complete geometry of
the TAVI prosthesis into the simulations is crucial. This would enable a more
accurate representation of the system and its behaviour. Secondly, exploring an
automated conversion process from STL to CAD formats could be beneficial to save
time, which should be considered as part of the computational workflow. Thirdly, it
is essential to study the multi-material behaviour of the solid domain. The observed
differences between the linear elastic and second-order Ogden models results suggest
that incorporating varying degrees of elasticity within the solid domain could lead
to a more robust model. This would contribute to a better understanding of the
system’s response and improve the overall accuracy of the simulations, as well as
their convergence. Fourthly, conducting a thorough parameter sensitivity analysis
is necessary to establish a reliable and precise FSI simulation setup that can effec-
tively model patient-specific boundary conditions. This step will ensure that the
simulations are robust and accurately capture the relevant factors influencing the
outcome. Finally, exploring alternative numerical methods could help address sim-
ulation convergence issues. By investigating different approaches, we can optimize
the simulations and improve their convergence and reliability.

In terms of computational resources, the availability of new tools and more pow-
erful computers can significantly reduce the simulation time, rendering computa-
tional resources and time constraints less problematic. This advancement would
allow for quicker analysis and facilitate more comprehensive investigations.

4.3 Conclusions
Mechanisms underlying thrombus formation after TAVI are still multi-factorial and
not well-understood. DRT is also one of the most challenging complications after
TAVI given that only patients who undergo CT angiography after TAVI can be di-
agnosed. In this context, early detection of patients who would eventually develop
DRT would help their treatment and follow-up.

In this study, we have introduced a workflow based on an FSI approach to assess


the risk of developing DRT post-TAVI. Through sensitivity analysis and calibration,
we tested an FSI simulation setup and obtained promising results that highlight the
advantages of this approach. Although FSI approach strengths have been demon-
strated, it is essential to validate the proposed workflow under physiological condi-
tions to assess thrombosis risk after TAVI allocation.

To achieve simulations that are more physiologically representative and patient-


specific, an exhaustive calibration of simulation parameters is still the primary step.
Additionally, exploring other numerical methods should be considered to enhance
the simulation convergence. However, the feasibility of including these steps in
future works should be evaluated based on computational resources.

30
Bibliography
[1] Salud y Medicina. Diez años desde la primera implantación de la TAVI en
España. url: https://www.saludymedicina.org/post/diez-anos-desde-
la-primera-implantacion-de-la-tavi-en-espana. (accessed: 09.12.2022).
[2] A. Íñiguez-Romo. “Outcomes of transcatheter aortic valve implantation in
Spain through the Activity Registry of Specialized Health Care”. In: REC:
Interventional Cardiology 4 (2022), pp. 123–131. doi: https://doi.org/10.
24875/RECICE.M21000259.
[3] S. G. Rouleau. “Transcatheter aortic valve replacement complications: A nar-
rative review for emergency clinicians.” In: The American Journal of Emer-
gency Medicine 56 (2022), pp. 77–86. doi: https://doi.org/10.1016/j.
ajem.2022.03.042.
[4] N. Ali. “A review of the technical challenges and optimal valve choices in
different clinical and anatomic scenarios”. In: 2019.
[5] P. Pasquale Leone. “Prosthesis Tailoring for Patients Undergoing Transcatheter
Aortic Valve Implantation”. In: JACC. Cardiovascular Imaging 12.1 (2023),
p. 338. doi: https://doi.org/10.3390/jcm12010338.
[6] P. Blanke. “Computed tomography imaging in the context of transcatheter aor-
tic valve implantation (TAVI)/transcatheter aortic valve replacement (TAVR):
An expert consensus document of the society of cardiovascular computed to-
mography”. In: JACC. Cardiovascular Imaging 12.1 (2019), pp. 1–24. doi:
https://doi.org/10.1016/j.jcmg.2018.12.003.
[7] A. Mayr. “Is MRI equivalent to CT in the guidance of TAVR? A pilot study”.
In: European Radiology 28.11 (2018), pp. 4625–4634. doi: https://doi.org/
10.1007/s00330-018-5386-2.
[8] A. Meyer. “Reliability and Influence on Decision Making of fully-automated
vs. semi-automated Software Packages for Procedural Planning in TAVI.” In:
Scientific Reports 10.1 (2020). doi: https://doi.org/10.1038/s41598-
020-67111-5.
[9] HeartNavigator. Insightful planning and guidance for Structural Heart Disease
Procedures. url: https : / / www . philips . com . au / healthcare / product /
HCOPT201/heartnavigator. (accessed: 28.01.2023).
[10] L. Biasco. “Access sites for TAVI: Patient selection criteria, technical aspects,
and outcomes”. In: Frontiers in Cardiovascular Medicine 5 (2018). doi: https:
//doi.org/10.3389/fcvm.2018.00088.
[11] A. Intorcia. “Management of transcatheter aortic valve implantation and com-
plex aorta anatomy: The importance of pre-procedural planning”. In: Interna-
tional Journal of Environmental Research and Public Health 19.8 (2022). doi:
https://doi.org/10.3390/ijerph19084763.

31
[12] S. Bhushan. “Paravalvular leak after transcatheter aortic valve implantation
its incidence, diagnosis, clinical implications, prevention, management, and fu-
ture perspectives: A review article”. In: Current Problems in Cardiology 47.10
(2022). doi: https://doi.org/10.1016/j.cpcardiol.2021.100957.
[13] S. Milhorini. “How valvular calcification can affect the outcomes of tran-
scatheter aortic valve implantation”. In: Expert Review of Medical Devices
17.8 (2020). doi: https://doi.org/10.1080/17434440.2020.1789456.
[14] R. Jabbour. “Delayed coronary occlusion after transcatheter aortic valve im-
plantation: Implications for new transcatheter heart valve design and patient
management”. In: Interventional Cardiology 13.3 (2018), pp. 137–139. doi:
https://doi.org/10.15420/icr.2018.24.2.
[15] P. MacCarthy. “Minimizing permanent pacemaker implantation (PPI) after
TAVI”. In: The British Journal of Cardiology 28.2 (2021), p. 20. doi: https:
//doi.org/10.5837/bjc.2021.020.
[16] L. Sondergaard. “Leaflet thrombosis after TAVI”. In: European Heart Journal
38.36 (2017), pp. 2702–2703. doi: https://doi.org/10.1093/eurheartj/
ehx473.
[17] M. Martín. “Transcatheter aortic valve implantation and subclinical and clin-
ical leaflet thrombosis: Multimodality imaging for diagnosis and risk stratifi-
cation”. In: European Cardiology 16.35 (2021). doi: https://doi.org/10.
15420/ecr.2021.09.
[18] M. Ranasinghe. “Thromboembolic and Bleeding Complications in Transcatheter
Aortic Valve Implantation: Insights on Mechanisms, Prophylaxis and Ther-
apy.” In: Journal of Clinical Medicine 8.2 (2019), p. 280. doi: https://doi.
org/10.3390/jcm8020280.
[19] B. Ncho. “In-Vitro Assessment of the Effects of Transcatheter Aortic Valve
Leaflet Design on Neo-Sinus Geometry and Flow”. In: Annals of Biomedical
Engineering 49.3 (2020), pp. 1046–1057. doi: https://doi.org/10.1007/
s10439-020-02664-0.
[20] Mathew N. Hindi. “Antithrombotic Therapy After Transcatheter Aortic Valve
Replacement: An Overview”. In: Structural Heart 6.5 (2022), p. 100085. doi:
https://doi.org/10.1016/j.shj.2022.100085.
[21] R. R. Makkar. “Possible subclinical leaflet thrombosis in bioprosthetic aortic
valves”. In: The New England Journal of Medicine 373.21 (2015), pp. 2015–
2024. doi: https://doi.org/10.1056/NEJMoa1509233.
[22] H. Hatoum. “Predictive model for thrombus formation after transcatheter
valve replacement”. In: Journal of the Royal Society, Interface 12.6 (2021),
pp. 576–588. doi: https://doi.org/10.1007/s13239-021-00596-x.
[23] V. Raghav. “Transcatheter aortic valve thrombosis: a review of potential mech-
anisms”. In: Journal of the Royal Society, Interface 18.184 (2021). doi: https:
//doi.org/10.1098/rsif.2021.0599.

32
[24] W. Mao. “Fluid-Structure Interaction Study of Transcatheter Aortic Valve
Dynamics Using Smoothed Particle Hydrodynamics.” In: Cardiovascular en-
gineering and technology 7.4 (2016), pp. 374–388. doi: https://doi.org/10.
1007/s13239-016-0285-7.
[25] H. S. Kandail. “Impact of annular and supra-annular CoreValve deployment
locations on aortic and coronary artery hemodynamics.” In: Journal of the
Mechanical Behavior of Biomedical Materials 86 (2018), pp. 131–142. doi:
https://doi.org/10.1016/j.jmbbm.2018.06.032.
[26] G. Luraghi. “On the modeling of patient-specific transcatheter aortic valve
replacement: A fluid-structure interaction approach.” In: Cardiovascular En-
gineering and Technology 10.3 (2019), pp. 437–455. doi: https://doi.org/
10.1007/s13239-019-00427-0.
[27] Adi A. Basri. “Fluid Structure Interaction on Paravalvular Leakage of Tran-
scatheter Aortic Valve Implantation Related to Aortic Stenosis: A Patient-
Specific Case”. In: Computational and Mathematical Methods in Medicine 2020
(2020), pp. 1–22. doi: https://doi.org/10.1155/2020/9163085.
[28] V. Govindarajan. “Improving transcatheter aortic valve interventional pre-
dictability via fluid-structure interaction modelling using patient-specific anatomy.”
In: Royal Society Open Science 9.2 (2022), pp. 512–515. doi: https://doi.
org/10.1098/rsos.211694.
[29] F. Borowski. “Validation of a Fluid Structure Interaction model for TAVR
using Particle Image Velocimetry.” In: Current Directions in Biomedical En-
gineering 8.2 (2022), pp. 512–515. doi: https://doi.org/10.1515/cdbme-
2022-1131.
[30] X. Liu. “Fluid-Structure Interaction Analysis on the Influence of the Aortic
Valve Stent Leaflet Structure in Hemodynamics.” In: Frontiers in physiology
13.6 (2022). doi: https://doi.org/10.3389/fphys.2022.904453.
[31] J. A. Brown. “Patient-specific immersed finite element-difference model of
transcatheter aortic valve replacement.” In: Annals of Biomedical Engineering
51.1 (2023), pp. 103–116. doi: https://doi.org/10.1007/s10439- 022-
03047-3.
[32] D. Oks. “Fluid-structure interaction analysis of eccentricity and leaflet rigidity
on thrombosis biomarkers in bioprosthetic aortic valve replacements”. In: In-
ternational Journal for Numerical Methods in Biomedical Engineering 38.12
(2022). doi: https://doi.org/10.1002/cnm.3649.
[33] Ram P. Ghosh. “Numerical evaluation of transcatheter aortic valve perfor-
mance during heart beating and its post-deployment fluid–structure interac-
tion analysis”. In: Biomechanics and Modeling in Mechanobiology 19.5 (2020),
pp. 1725–1740. doi: https://doi.org/10.1007/s10237-020-01304-9.
[34] M. S. Sadrabadi. “Fluid-structure coupled biotransport processes in aortic
valve disease”. In: Journal of Biomechanics 117 (2021), p. 110239. doi: https:
//doi.org/10.1016/j.jbiomech.2021.110239.

33
[35] K. Vahidkhah. “Blood Stasis on Transcatheter Valve Leaflets and Implications
for Valve-in-Valve Leaflet Thrombosis”. In: The Annals of Thoracic Surgery
104.3 (2017), pp. 751–759. doi: https://doi.org/10.1016/j.athoracsur.
2017.02.052.
[36] I. Fumagalli. “Fluid-structure interaction analysis of transcatheter aortic valve
implantation”. In: International Journal for Numerical Methods in Biomedical
Engineering (2023). doi: https://doi.org/10.1002/cnm.3704.
[37] N. Pil. “Influence of Aortic Valve Leaflet Material Model on Hemodynamic
Features in Healthy and Pathological States”. In: Mathematics 11.2 (2023),
p. 428. doi: https://doi.org/10.3390/math11020428.
[38] A. R. Prisco. “The native aortic valve reduces paravalvular leak in TAVR
patients.” In: Frontiers in Physiology 13 (2022). doi: https://doi.org/10.
3389/fphys.2022.910016.
[39] A. K. Anidis. Aortic flow patterns after simulated implantation of transcatheter
aortic valves. url: https://efstathopoulos.org/wp- content/uploads/
2017 / 03 / 79 _ hjc _ 2015 _ kopanidis _ aortic - flow - patterns - after -
simulated - implantation - of - transcatheter - aortic - valves - 1 . pdf.
(accessed: 29.01.2023).
[40] J. Petersen. “Comparison of aortic flow patterns before and after transcatheter
aortic valve implantation.” In: Cardiovascular Engineering and Technology 3.1
(2012), pp. 123–135. doi: https://doi.org/10.1007/s13239-011-0073-3.
[41] D. Bluestein. “Research approaches for studying flow-induced thromboembolic
complications in blood recirculating devices”. In: Expert Review of Medical
Devices 1.1 (2004), pp. 65–80. doi: https://doi.org/10.1586/17434440.
1.1.65.
[42] A. Yazdani. “A General Shear-Dependent Model for Thrombus Formation”.
In: PLOS Computational Biology 13.1 (2017). doi: https://doi.org/10.
1371/journal.pcbi.1005291.
[43] Y. Xuan. “Stent and leaflet stresses in a 26-mm first-generation balloon-expandable
transcatheter aortic valve”. In: The Journal of Thoracic and Cardiovascular
Surgery 153.5 (2017), pp. 1065–1073. doi: https://doi.org/10.1016/j.
jtcvs.2016.12.016.
[44] M. Hsu. “Fluid–structure interaction analysis of bioprosthetic heart valves:
significance of arterial wall deformation”. In: Computational Mechanics 54.4
(2014), pp. 1055–1071. doi: https://doi.org/10.1007/s00466-014-1059-
4.
[45] S. Barati. “A computational optimization study of a self-expandable tran-
scatheter aortic valve”. In: Computers in Biology and Medicine 139 (2021),
p. 104942. doi: https://doi.org/10.1016/j.compbiomed.2021.104942.
[46] Zachary Tugaoen et al. “The selection of transcatheter heart valves in tran-
scatheter aortic valve replacement”. In: Trends in Cardiovascular Medicine
32.8 (2022), pp. 513–522. doi: https://doi.org/10.1016/j.tcm.2021.10.
002.

34
[47] J. E. Davies. “Attenuation of Wave Reflection by Wave Entrapment Creates a
“Horizon Effect” in the Human Aorta”. In: Hypertension 60.3 (2012), pp. 778–
785. doi: https://doi.org/10.1161/hypertensionaha.111.180604.
[48] T. Christian Gasser, Ray W Ogden, and Gerhard A Holzapfel. “Hyperelastic
modelling of arterial layers with distributed collagen fibre orientations”. In:
Journal of The Royal Society Interface 3.6 (2005), pp. 15–35. doi: https :
//doi.org/10.1098/rsif.2005.0073.
[49] Gerhard A. Holzapfel, Thomas C. Gasser, and Ray W. Ogden. “A New Con-
stitutive Framework for Arterial Wall Mechanics and a Comparative Study of
Material Models”. In: Journal of The Royal Society Interface 61.1/3 (2000),
pp. 1–48. doi: https://doi.org/10.1023/a:1010835316564.
[50] S. A. Berger. “Flows in Stenotic Vessels”. In: Annual Review of Fluid Mechanics
32.1 (2000), pp. 347–382. doi: https://doi.org/10.1146/annurev.fluid.
32.1.347.
[51] Rammohan Subramania Raja. “Coupled fluid structure interaction analysis on
a cylinder exposed to ocean wave loading”. In: (2012).
[52] Mikel Ezkurra. “Analysis of One-Way and Two-Way FSI Approaches to Char-
acterise the Flow Regime and the Mechanical Behaviour during Closing Ma-
noeuvring Operation of a Butterfly Valve”. In: World Academy of Science,
Engineering and Technology, International Journal of Mechanical and Mate-
rials Engineering 12 (2018), pp. 313–319. doi: http://doi.org/10.5281/
zenodo.1316365.
[53] S. K. Chimakurthi. “ANSYS Workbench System Coupling: a state-of-the-art
computational framework for analyzing multiphysics problems”. In: Engineer-
ing with Computers 34.2 (2017), pp. 385–411. doi: https://doi.org/10.
1007/s00366-017-0548-4.
[54] D. Pott. “Hemodynamics inside the neo- and native sinus after TAVR: Effects
of implant depth and cardiac output on flow field and coronary flow”. In:
Artificial Organs 45.1 (2020), pp. 68–78. doi: https://doi.org/10.1111/
aor.13789.
[55] H. Hatoum. “Impact of patient-specific morphologies on sinus flow stasis in
transcatheter aortic valve replacement: An in vitro study”. In: The Journal of
Thoracic and Cardiovascular Surgery 157.2 (2019), pp. 540–549. doi: https:
//doi.org/10.1016/j.jtcvs.2018.05.086.
[56] S. Barati. “Patient-specific multi-scale design optimization of transcatheter
aortic valve stents”. In: Computer Methods and Programs in Biomedicine 221
(2022), p. 106912. doi: https://doi.org/10.1016/j.cmpb.2022.106912.
[57] A. Schiavone. “A study of balloon type, system constraint and artery consti-
tutive model used in finite element simulation of stent deployment”. In: Me-
chanics of Advanced Materials and Modern Processes 1.1 (2015). doi: https:
//doi.org/10.1186/s40759-014-0002-x.

35
A Appendix
A.1 TAVI prosthesis CAD

Figure A1: Metallic frame component designed in SolidWorks from literature measure-
ments and drawings [45, 56].

Figure A2: Metallic frame designed in SolidWorks inspired on TAVI frames design.

36
A.2 Boundary conditions
This section is based on the ongoing Master’s Thesis of Laura Escot.
The curve profile was extracted using WebPlotDigitalizer (California, USA). Sub-
sequently, MATLAB (MathWorks, California, USA) was employed to post-process
the segmented curve. This involved discretizing the velocity data at intervals of
0.001s for a complete patient-specific beat lasting 0.882s. The starting point was set
at the beginning of the ventricular diastole phase.

Figure A3: (a) LVOT Doppler echocardiography of the patient, and (b) the velocity curve
segmented and post-processed to be adjusted and synchronized to the patient-specific beat
(0.882s per beat). LVOT = left ventricular outflow tract.

37
A.3 Mesh convergence study

Figure A4: Solid domain mesh with element sizes ranging from 1 - 4 mm, displaying
surface discontinuities in the leaflets, causing a deviation from the original CAD geometry.
CAD = computer-aided design.

38
A.4 Ogden model results
This section presents the results of the FSI simulation using the Ogden model instead
of linear elastic. The simulation setup consisted of the following components:

1. Meshing element size ranging from 1 to 3 mm.

2. Inlet: PS velocity profile.

3. Outlet: constant pressure of 100 Pa.

4. Aortic walls and leaflets described by a second-order Ogden model (Table A1).

5. Under-relaxation factor set to 0.45.

ρ (kg/mm3 ) µ1 µ2 µ3 α1 α2 α3 D1
1.07e−6 -4.73 1.70 3.09 -0.39 4.41 -3.25 3.63e−6

Table A1: Values of the Ogden model parameters for modelling homogeneous artery [57].

Due to stability issues, the simulation could not be fully completed. However,
results were obtained for the initial 0.25s. Figure A5 shows the obtained results
at four specific instances. While the leaflets did not experience any mechanical
deformations, the aortic walls within the fluid domain exhibited deformations as
a result of the mechanical properties defined by the second-order Ogden model in
the solid domain. It is worth noting that this was the only conducted test where
significant deformation was observed in the aortic walls.

Figure A5: FSI-derived blood flow velocity streamlines at four instances. The top row
(a-d) shows the leaflets opening and closing kinematics at the corresponding instances. FSI
= fluid-structure interaction.

39
A.5 Von Misses stress and shear stress

Figure A6: Leaflet kinematics FSI-derived von Mises stress (MPa) at six instances during
leaflet opening and closing. FSI = fluid-structure interaction.

Figure A7: Leaflet kinematics FSI-derived shear stress (MPa) at six instances during
leaflet opening and closing. FSI = fluid-structure interaction.

40

You might also like