You are on page 1of 41

Journal Pre-proof

Computational analysis of Nitinol stent-graft for endovascular


aortic repair (EVAR) of abdominal aortic aneurysm (AAA):
Crimping, sealing and fluid-structure interaction (FSI)

Raja Jayendiran, Bakr Nour, Annie Ruimi

PII: S0167-5273(19)31364-6
DOI: https://doi.org/10.1016/j.ijcard.2019.11.091
Reference: IJCA 28124

To appear in: International Journal of Cardiology

Received date: 13 March 2019


Revised date: 20 June 2019
Accepted date: 8 November 2019

Please cite this article as: R. Jayendiran, B. Nour and A. Ruimi, Computational analysis
of Nitinol stent-graft for endovascular aortic repair (EVAR) of abdominal aortic aneurysm
(AAA): Crimping, sealing and fluid-structure interaction (FSI), International Journal of
Cardiology(2019), https://doi.org/10.1016/j.ijcard.2019.11.091

This is a PDF file of an article that has undergone enhancements after acceptance, such
as the addition of a cover page and metadata, and formatting for readability, but it is
not yet the definitive version of record. This version will undergo additional copyediting,
typesetting and review before it is published in its final form, but we are providing this
version to give early visibility of the article. Please note that, during the production
process, errors may be discovered which could affect the content, and all legal disclaimers
that apply to the journal pertain.

© 2019 Published by Elsevier.


Journal Pre-proof

Computational analysis of Nitinol stent-graft for


endovascular aortic repair (EVAR) of abdominal
aortic aneurysm (AAA): crimping, sealing and
fluid-structure interaction (FSI)

Raja Jayendiran, Ph.D.1 ∗ Bakr Nour, MD2 and


Annie Ruimi, Ph.D.1

of
1 Mechanical Engineering Program, Texas A&M University at Qatar, Doha (Qatar)

ro
2 Weill Cornell Medicine-Qatar, Doha (Qatar)
-p
Abstract
re

Objectives: We evaluate the crimping strain, sealing stress and contact forces on a Niti-
lP

nol stent deployed in the aorta during endovascular aortic (or aneurysm) repair (EVAR)
na

procedures. Nitinol shape memory effect (SME) is used. We also study the fluid-structure

interaction (FSI) of the blood flow on the stented aorta.


ur

Methods: We employ Solidworks to generate a closed-cell honeycomb stent structure


Jo

used to treat abdominal aortic aneurysm (AAA).We use the commercial Abaqus/Simulia

finite element (FEM) simulation package to study the displacements and stresses experi-

enced by the stent during the crimping phase and deployment into the aortic segment. The

Nitinol stent is covered with Dacron, a popular graft material. We implement our own user-

material (UMAT) subroutines to model the shape memory effect (SME) of Nitinol. The

effect of the stent geometry is analyzed. We use the FSI analysis in Abaqus/Simulia to

understand the effect of hemodynamic loading on the stent.

Results: Results indicate that the crimping strain increases as the stent strut spacing

Preprint submitted to Elsevier 24 June 2019


Journal Pre-proof

increases. This is also the case for the radius of curvature. Maximum strains developed on

the stent during crimping are in the order of 10%. Stresses exerted by the stent needed to

completely seal the aorta are found to be below the yield stress values of Nitinol (700MPa).

Wall shear stresses (WSS) on the stented aorta are close to WSS obtained on the aorta

alone.

Conclusion: Using Nitinol’s thermo-reactivity property as opposed to its superelasticity

causes the stent-graft to deploy more gently.

of
ro
Key words: Stent-graft, Crimping, Sealing stress, Fluid-Structure Interaction, Wall shear
stress, Nitinol -p
re

1 Background
lP
na

Aortic aneurysms refer to a localized widening of the aorta to at least 1.5 times its original size.

The most common is for the aneurysm to occur in the part of the aorta that passes through the ab-
ur

domen (abdominal aortic aneurysm AAA), but it can also be located in the chest cavity (thoracic
Jo

aortic aneurysm TAA). Treatments vary depending on the patient’s condition and on a case-to-case

basis. Choices include open-surgery procedures which are largely invasive or endovascular aortic (or

aneurysm) repair (EVAR). These latter involve the insertion into the artery of a metallic scaffold

(stent), sometimes lined or covered with a fabric (graft), via incisions in the leg rather than in the

abdomen. Because of that, EVAR are known to result in less trauma, earlier return to daily activities,

reduced mortality and lower morbidity [1, 2].


∗ Corresponding author.
Email addresses: jayendiran.raja@qatar.tamu.edu (Raja Jayendiran, Ph.D.1 ),
annie.ruimi@qatar.tamu.edu (Annie Ruimi, Ph.D.1 ).

2
Journal Pre-proof

The stent primary function is to reinforce the aortic walls and to prevent the blood from flowing into

the aneurysm sac [3, 4]. Various stent designs have appeared on the market over the last 20 years but

self-expendable metallic stents such as those made of shape memory alloys (SMAs) have seen a large

increase in popularity [5, 6].

Among SMAs, Nitinol (NiTi) is an alloy of nickel and titanium. A temperature change can trans-

form its crystal structure from an austenite form at high temperature to a martensite form (at low

of
temperature). This is known as shape-memory effect (SME). In addition, Nitinol possesses a supere-

ro
lastic property, i.e. the ability for the material to regain its original shape when removing the loads,
-p
alike a ’super-spring’ [7]. In such instances, the material can stretch up to 10 times its original size.
re
Nitinol is also known for its superior biocompatibility and good resistance to corrosion and fatigue
lP

[8, 9, 10]. Among graft materials, the most popular are polytetrafluoroethylene (PTFE) and polyethy-

lene terephthalate (PET). PTFE is a very flexible synthetic polymer that can be formed into complex
na

three-dimensional shapes. PET (tradename Dacron) is a polyester fabric derived from ethylene glycol
ur

and terephthalic acid. It is a very rigid plastic that needs to be woven to create a flexible graft [11].
Jo

The stent-graft needs to be sufficiently compressed to fit into a thin catheter before being released into

the artery. During the compressing process, strains are generated. Those strains referred as crimping

strains measure the change in shape and size and must remain below a certain value for fracture not

to occur [12, 13]. In the case of Nitinol, the fracture criterion is strain-based as opposed to other

materials for which fracture criterion is stress-based [14].

Once the stent has been placed into the artery, it can expand and recover its initial shape and size. The

stresses generated during the expansion phase must be large enough to allow the stent to adhere and

3
Journal Pre-proof

completely seal the artery (sealing stress). Otherwise, the stent might migrate and let the blood leak

(endoleak), an undesirable occurrence [15, 16]. However, too high of a stress on the aortic walls may

not be desirable either [17, 18].

The aorta is a very flexible structure. Its walls deform greatly during a typical heart cycle and can

result in a change in diameter of more than 10% [19]. It is therefore important to understand how

the blood flow affects the mechanical behavior of the stented aortic segment once the stent has fully

of
expanded. This can be done with a fluid-structure interaction analysis (FSI). This type on analysis rests

ro
on the assumption that the walls are not rigid. The method combines a computational fluid dynamics
-p
analysis (CFD) for the fluid to a finite element analysis (FEA/FEM) for the solid. Such computational
re
analyses are valid alternatives to experimental studies on aorta or graft which are possible but difficult
lP

to achieve [20, 21]. Earlier studies (see for example [22, 23]) have reported results on stent-graft

deployment using a FEM approach but without taking into account the interaction of the fluid on the
na

structure.
ur

In this paper, we use the commercial finite element simulation package Abaqus/Simulia (Version
Jo

6.13) to evaluate the strains and stresses experienced by the stent during the crimping and deploy-

ment phases. The stent we design resembles a honeycomb structure with open-cells. We implement

our own user-material (UMat) subroutine to model the shape-memory effect of Nitinol. We identify

regions of the stent where strains are maximum. We proceed to a FSI analysis, -a built-in capability

in Abaqus/Simulia-, to evaluate radial displacements, von Mises stresses and wall shear stress (WSS)

on a stented aortic segment placed in an aneurysm.

We follow in part the analysis presented by Kleinstreuer et al. in [24] with the following modifi-

4
Journal Pre-proof

cations: i) we use a closed-cell honeycomb stent structure when in [24], the stent is an open-cell

diamond design and ii) we assume that thermoreactivity of Nitinol is responsible for the stent expan-

sion while in [24], removing the mechanical loads (i.e. Nitinol superelasticity property) causes the

stent deployment.

Details on the methodology including the material models used for the simulations are described in

Section 2 with the numerical schemes following in Section 3. Results are presented in Section 4 before

of
ending with some conclusive comments in Section 5.

ro
-p
2 Methodology
re
lP

2.1 Geometry
na

Aorta: The geometry we choose for the abdominal aortic segment is a simple cylindrical shell struc-
ur

ture, 12.6 cm long, 30 mm diameter and 2.61 mm thick in total. The portion of the aorta with an
Jo

aneurysm has a 50 mm diameter representing a 1.7 times dilatation compared to its original size. Pre-

vious work by the authors on using a patient-specific 3D reconstructed aortic anatomy have already

been published [25]. The three layers of the aortic walls (intima, media and adventitia) are taken into

consideration. Details regarding the physiology or structure of the human aorta can be found in [26].

The aorta is meshed with 24422 S4R (linear quadrilateral) shell elements and 24551 nodes. Shell ele-

ments are chosen because the aortic layers are thin (between 1/3 mm and 1 mm). Such elements allow

the modeling of thin features with much fewer elements than solid elements so the computational

time, especially during FSI, can be reduced. Shell elements are also easier to mesh and less prone to

5
Journal Pre-proof

returning negative Jacobian errors which might occur when using extremely thin solid layers [27].

Blood domain: The basis for FSI simulations is to insert a fluid domain into a solid domain of the

same dimensions (i.e. the aortic segment with dimensions noted above), provided conditions at the

interface between the two domains are enforced. Details follow in Section 2.3.

Stent: We use the same honeycomb stent structure we previously generated for the simulations re-

of
ported in [28], based on the design procedure outlined in [29]. In [28], we employed the ’wrap-

technique’ of the 2015 × 64 Edition of Solidworks to repeat a unit-cell geometry and construct a

ro
stent, 6 cm long, 3 cm in (outer) diameter comprising 12 rows of 8 struts, each 0.3 mm thick with seg-
-p
ments connecting the struts (i.e closed-cells). The resulting geometry is exported to Abaqus/Simulia
re
for analysis. The stent is meshed with 29995 linear triangular shell elements of type S3 and 21809
lP

nodes. The rationale for using shell elements is the same as the one discussed above for the aorta.
na

Graft: As in [28], Dacron is selected as the graft material and is chosen to cover the stent. It is

modelled as a cylindrical shell, 6 cm long, 3 cm in diameter and 0.361 mm thick. It is meshed into
ur

19840 linear quadrilateral (S4R) shell elements with 20000 nodes.


Jo

2.2 Material models

Aorta and Dacron: The aorta and the Dacron are represented with hyperelastic models. For the aorta,

we use the Holzapfel-Gasser-Ogden (HGO) model which was developed to model the anisotropy of

biological tissues, including compressibility as well as the randomness of the fibers distribution [30].

Details regarding the HGO model can be found in [31] or [32]. The Dacron is modeled as a neo-

Hookean material, a simpler hyperelastic model used when isotropy is assumed [30]. We nonetheless

6
Journal Pre-proof

realize that an anisotropic model may be more accurate to represent the corrugations of the Dacron

fabric [33]. Computational studies by the authors are underway using an anisotropic model for the

Dacron graft. Material properties were taken from [31, 34, 35].

Blood: The assumptions for the blood flow characteristics is that it is pulsatile, Newtonian, incom-

pressible and turbulent. These are chosen to represent realistic physiological conditions. The Newto-

nian fluid assumption is due to a high shear rate in the aorta [36, 37, 38]. Incompressibility is assumed

of
because the blood density varies less than 5%. Fully developed and turbulent flow is the result of im-

ro
perfections present on the aortic walls [39]. Additional parameters needed for the simulations include
-p
the blood density (1050 kg/m3 ) and a dynamic viscosity of 0.003 Pa.s. Details can be found in [32].
re
Stent: The stent we simulate is made of Nitinol, which has been described above as possessing both
lP

superelasticity and shape memory effect (SME) properties. Several models have been proposed in the

literature to describe SMAs. Here we use the model proposed by Lagoudas et al. in [40] which can
na

capture a linear thermoelastic response (up to 10% strain) and a non-linear transformation-hardening
ur

behavior for SMAs [40, 41]. However, Abaqus/Simulia does not have in its library of materials,
Jo

models that can describe SME so we had to implement our own UMAT (User-Material) subroutines.

2.3 Boundary conditions

Blood & stent: At the interface between the outer surface of the blood domain and the inner surface of

the solid domain (stent), conditions apply. The conditions are continuity of displacement and continu-

ity of the normal component of stress between the two domains. The tangential components of stress

(i.e. shear stresses) between the fluid and the solid are not equal. In addition, the no-slip condition

7
Journal Pre-proof

(i.e. condition of continuity of velocity at the interface between the two domains) is also enforced to

prevent the fluid from escaping through the solid domain.

Stent & graft: In our proposed design, the graft is used as a jacket that covers the stent. The artery is

constrained in the proximal and distal sections (i.e. the two end sections) to prevent rotations while

between them, in the middle section, the stent-graft can only experience radial displacements.

of
Graft & aorta: In this study, the pressure is given as input at both the inlet and outlet of the aortic

segment. Usually, Fourier series are used to represent the pressure profile during the heart beat as a

ro
time-dependent function (see for instance Eq. 1 ref [32]). Here, we use the pressure profile employed
-p
in [25].
re
lP

3 Numerical simulation scheme


na

3.1 Contact algorithm


ur
Jo

The domain under consideration involves various surfaces of contact needing a contact algorithm.

This is done using the movable rigid surface subroutine available in Abaqus/Simulia. In the subrou-

tine, the stent is deformable; the bottom plate is rigid and fixed while the top plate is rigid but movable

to allow the stent to compress. The top plate is the master surface and the stent, the slave surface. In

addition, a penalty interaction scheme is applied to the surfaces to render them impermeable and to

suppress any vibration resulting from the interaction of the stent and the artery. A contact pressure is

generated by the crimping tool (i.e. the movable rigid surface) with the assumption that no friction

exists between the tool and the stent. Additional details about the contact pressure can be found in

8
Journal Pre-proof

[24]. The crimping and expansion simulations are done on Abaqus/Standard (v 6.13-1) with a 1s time

step, taking into account the large deformations.

3.2 Cosimulation scheme

of
ro
We use the fluid-structure interaction (FSI) built-in capability in Abaqus/Simulia [42]to solve the
-p
equations representing the problem at hand. At the interface between the solid and the fluid domains,

data is exchanged with the cosimulation engine (CSE). Here, a dynamic implicit iterative technique
re

allows the exchange of fields between the domains for multiple times until an overall equilibrium
lP

is achieved at the current step. The pressure and wall shear stress values from the fluid surface are
na

transformed into an equivalent traction vector on the solid surface while the displacement and velocity

components of the solid surface are transformed into surface nodes (i.e. meshing). The scheme is
ur

illustrated in Fig. S1 a (refer Supplementary Figures). The Gauss-Seidel algorithm is a sequential


Jo

explicit coupling scheme used to obtain the field output data at a user-specified frequency (Fig. S1

b). The time duration to exchange the data between the FEA and CFD was set to 10−3 s for a total

time step of three seconds. Both FEA and CFD use the initial boundary conditions at the interface to

solve for values at the first time increment (i). At the next time increment (i+1), FEA uses the CFD

pressure solution pi at time (i) obtained from CFD to calculate the traction vectors. The FEA solver

updates the displacement and velocity values of the nodes ui which are then sent to CFD to solve for

the pressure and shear values at (i+1). A very small time increment in the order of 10−5 s was used in

the CFD such as to return a minimum residual.

9
Journal Pre-proof

4 Results

4.1 Model Validation

of
ro
-p
Experimental validation of the model we propose would be useful but very difficult to achieve. We

were however able to ascertain the validity of our model by comparing our results to results pub-
re

lished in the literature. In a 2008 study [24], the authors conduct a crimping analysis of a tubular,
lP

diamond-shape (open-cell) Nitinol stent. We ran a crimping analysis on a stent of the same geometry
na

and dimensions reported in the study: a cylinder 10 cm long and 3 cm in diameter, 0.28 mm thick

with 200 struts (10 rows of 20 struts) (refer Supplementary Figures, Fig. S2 a & b). The geometry
ur

was reproduced with Solidworks. The stent was constrained in the axial direction but allowed to de-
Jo

form freely in the radial direction. In the circumferential direction, symmetry was enforced. Radial

displacements were applied to the stent to compress it into a 0.8 cm diameter to fit into a catheter, as in

[24]. Our simulations returned a maximum 9% crimping strain on the stent (refer Supplementary Fig-

ures, Fig. S2 c) which occurred at the apex of the diamond cells, where the curvature changes sharply.

This value was also well below the critical 12% strain mentioned in [24] for Nitinol. In addition, the

contours obtained from our simulations were very close to those shown in [24]. Thus, we inferred that

the model we used for the diamond-shape stent was valid and that it could be manipulated to represent

other materials and other stent geometries such as the honeycomb stent we designed.

10
Journal Pre-proof

4.2 Crimping

We proceeded to the crimping analysis on the honeycomb stent-graft. The original 3 cm outer-

diameter needed to be reduced to a 0.8 cm diameter to fit into a 24F size guidewire, per the stent

manufacturers guidelines (1F = 0.333 mm). In order to do that and as described in [24], the stent

was constrained in the axial direction but allowed to deform freely in the radial direction. In the cir-

of
cumferential direction, symmetry was enforced. Crimping was done at ambient conditions, i.e. a 293

ro
K (≈ 20o C) temperature, when Nitinol is in its austenite phase. Radial displacements were applied
-p
to compress the stent until it was completely crimped, at which point, Nitinol was in its martensite
re
phase. Nitinol phase transformation diagram is shown in Fig. S3 (refer Supplementary Figures).
lP

We then examined how the strut geometry affected the crimping strain values. The maximum crimp-
na

ing strains experienced by the stent are shown on Fig. 1 for various strut curvatures (Rc ) and strut

spacing (H). Values indicate that the crimping strain increases linearly as the strut spacing increases.
ur

Doubling the strut spacing (H = 12 mm vs. H = 6 mm) will result in a ≈ 7% higher crimping strain
Jo

(8.82% vs. 9.44%). Likewise, the crimping strain increases linearly as the strut curvature increases.

However, a design trade-off is that the higher the radius of curvature, the more potential for rupture.

On the other hand, decreasing the strut spacing will increase the stent stiffness but as more material is

added, biocompatibility may become an issue. Simulations show that the maximum crimping strain

is obtained for H = 12 mm and Rc = 1.5 mm and is equal to 9.45%. It also shows that the maximum

crimping strain is experienced at the stent curvatures. This maximum strain is still below the critical

strain (i.e. 12%) of Nitinol reported in [24]. That means that plastic deformations will not occur dur-

ing the crimping process and that the stent will have the ability to regain its original shape once it is

11
Journal Pre-proof

deployed.

4.3 Stent release

To simulate the stent release from the thin catheter into the diseased portion of the aorta (i.e. aorta

with aneurysm), we ’reversed’ the crimping process and applied outward radial displacements to the

of
stent until we reached an unloaded configuration, i.e. a state of zero stress (Fig.2). This was done

while maintaining a 293 K temperature when Nitinol is in martensite phase.

ro
-p
When the stent is crimped, the maximum strain is 9.5% (Fig. 1, for H = 12 mm and Rc = 1.5 mm). As

the stent is released, the stent diameter increases. This is due to some elastic deformation as reported
re

in [43]. When the loads are removed, there is still a 6.5% residual strain (Fig. 2) on the stent that need
lP

to be removed for the stent to return to its original size. These strains are residual strains which have
na

accumulated during the crimping process. In order to remove these residual strains, we increased the

temperature from 293 K to 310 K (i.e. body temperature) and the stent returned to its original 30 mm
ur

diameter. After the stent has been released, the temperature is further increased from 310 K to 316
Jo

K, which corresponds to the final austenite temperature. At that temperature, there are no residual

strains.

4.4 Expansion and Sealing

For the expansion and sealing simulations of the stent, the stent diameter is still at 30 mm, but the

diameter of the aorta is set at 27 mm. This ’oversized’ stent value is chosen so that the stent expe-

riences a 11% increase in size (from 27 mm at insertion into the aorta to 30 mm in its expanded

12
Journal Pre-proof

state) as described in [44]. Once the stent has expanded, it can totally adhere (seal) to the aortic walls

and ’pushes’ against them. At the region of contact, both the stent and the aorta have now the same

diameter (30 mm).

Figure 2 shows the von Misses stresses contours recorded on the stented aorta, first during initial

contact between the stent and the aorta and next, when the stent has fully expanded (i.e. the aorta

of
is completely sealed). In both cases, the maximum stress (0.3 MPa) occurs at both ends of the stent

ro
and decreases the further away from the stent. During the initial contact, a large portion (≈ 3/4th) of

the stented area is under zero stress (blue). Eventually, stress builds-up again and when the stent has
-p
completely sealed the aorta, the stress is zero only at the center of the stented aortic region. In the
re
deployed state, the stent is in austenite phase and the 0.3 MPa recorded value of stress is below the
lP

yield stress of Nitinol (between 195-700 MPa), in austenite phase as reported in [45]. The low stress

values are desirable as high stress on the aortic walls lead to the proliferation of smooth muscle cells
na

(restenosis) [13] and may also damage the vessels endothelial cells, triggering pressure necrosis [24]
ur
Jo

Simulations were conducted for various stent diameters keeping the diameter of the aorta constant

and values of contact forces between the outer stent surface and the inner surface of the aorta were

extracted from the simulations. They are shown in Fig. 3. Values indicate that after deployment, the

contact force between the stent and the aortic walls increases linearly with increase in stent diameter.

When the stent expands from 27 mm to 30 mm, the contact force is about 5.82 N. Reference [46]

reports that this force is large enough to keep the stent in place and that migration is unlikely to occur.

It is important to remember that a high contact force may cause damage to the vessel walls but at the

same time, a smaller contact force may mean that the stent is not properly anchored.

13
Journal Pre-proof

4.5 Fluid-structure interaction

As alluded in the introduction, the aorta is a very flexible and deformable structure for which dis-

placements have been recorded to reach up to 10% during a cardiac cycle [19]. It is thus important to

evaluate the effect that the blood flow has on the stented aorta. Details regarding the FSI methodology

leading to the evaluation of radial displacements, von Mises stress and WSS on the graft, stent and

of
aortic walls due to the blood flow have been validated by the authors and presented in [25, 28, 32].

ro
Figure 4 shows the blood pressure profile used as input in the simulations. The pressure oscillates

between 110 mmHg/75 mmHg (millimeter of mercury) where 110 mmHg represents the systolic
-p
(peak) pressure and 75 mmHg represents the diastolic (lowest) pressure, typically of the blood pres-
re
sure fluctuation in an adult. This pressure profile has been previously compared to a pressure profile
lP

obtained experimentally by Di Martino et al. [47] using Doppler measurements. Here, the temperature

is maintained at 310 K when Nitinol is in its austenite phase.


na

For the FSI simulations, the healthy portion of the aorta is taken as having a 30 mm diameter, while
ur

the aneurysm has a 50 mm diameter (at maximum).


Jo

One can see from Fig. 4, that far away from the aneurysm sac, the aortic walls displace 1.4 mm radially

(red) while the stresses recorded on them approaches 0.66 MPa (light green). The stent-graft expe-

riences only 0.6 mm displacement (light green) and the aneurysm sac has negligible displacements

(light and dark blue).

The maximum von Mises stress recorded on the stent struts is 205 MPa (red). This value is acceptable

as it is below the yield stress of Nitinol in austenite phase [45]. The maximum stress generated on

the Dacron is 0.62 MPa (bright green). This is below but close to the von Mises stress of the aortic

14
Journal Pre-proof

walls (0.67 MPa) [24]. At the most dilated location of the aneurysm, the displacement is negligible

and the stress is 14.8 Pa (blue). This shows that when the stent-graft is fully deployed, the effect of

the hemodynamic load on it is negligible and that the stented aorta is unlikely to rupture.

Figure 4 on the right, also shows that the wall shear stress (WSS) attains a maximum value of 4.21

Pa (red), far away from the aneurysm sac. The wall shear stresses are quantities only relevant to the

blood domain (depicted in a red cylindrical domain). In the stented aortic region (brown), the value is

of
4.19 Pa i.e. there is almost no change from that in the healthy portion.

ro
In the recent literature [48], the average WSS in the descending aorta of a healthy patient is reported
-p
as 2.23±1.04 Pa. This value is obtained using a 4D flow magnetic resonance image (MRI). The value
re
we found is about double. A reason for that may be that we used an idealized velocity map (uniform
lP

flow in the lumen) while it is probably not the case in reality. Another reason is that our WSS value is

obtained assuming a stented aorta when in [48], the value refers to the aorta alone.
na
ur

4.6 Discussion
Jo

Designing stent involves selecting a fine combination of material and geometry capable of sustaining

various changes of configurations during the crimping, expansion and sealing phases. Among the

sought-after characteristics of stents materials, they should be resistant to corrosion, possess good or

superior biocompatibility and have favorable radiographic properties. The material chosen, Nitinol

is resistant to corrosion due to its titanium content, and has also been shown to exhibit excellent

biocompatibility. It is true that it has a limited fluoroscopic visualization but on the other hand, it can

produce excellent MRIs results [49].

15
Journal Pre-proof

To decide on an optimal stent geometry, we simulated stents with various strut spacing and radius of

curvature. We showed that the maximum strain obtained (9.45%) was below 12%, the value of the

critical strain for Nitinol reported in [24]. We also showed that our honeycomb close-cell stent design

did not suffer any plastic deformations during the crimping process.

Choosing the strut thickness is also an important step in the design process as thinner struts have been

of
shown to decrease the occurrence of restenosis, one of the main complication of post-stent deployment

ro
[50]. In addition, the hemodynamic load can result in radial buckling and proper strut thickness can

help in preventing that. In this design, we choose a 0.3 mm strut thickness and the simulations showed
-p
that the stent was able to maintain its size and shape when subjected to the blood flow, and helping
re
the blood to flow smoothly at the same time. This 0.3 mm value was a good choice to prevent the
lP

stent from buckling [10]. On the other hand, this thickness may be considered high which may induce

restenosis. Additional study on the topic is needed.


na
ur

In regards to device fracture and failure, regions of high stress should be minimized. Our simulation
Jo

results showed that the maximum stress on the stent due to the blood flow was well below the yield

stress of Nitinol. In addition, the sealing analysis showed that the stent was capable of withstanding

the compressive load exerted by the arterial walls. The maximum stress at the interface between the

stent and the artery was recorded to be 0.3 MPa and the contact force was calculated at 5.82 N, enough

for the stent to be completely sealed to the aortic walls without the stent changing shape. The contact

force was also strong enough to prevent the stent from migrating and the blood to leak. In this study,

the wall shear stress (WSS) values were shown to be between 1-7 Pa, the same as those found in a

healthy vessel. Stent restenosis are linked to high WSS values.

16
Journal Pre-proof

The stent design we proposed saw a 26% reduction in diameter (i.e. from 30 mm to 8 mm) during

crimping allowing the stent fit properly into the catheter. Both, diameter and length of the stent were

completely recovered after expansion.

Owing to all the reasons above, the computational simulations results point to the stent design we pro-

posed as an excellent candidate for EVAR procedures. Long term clinical trials will need to validate

this further.

of
ro
4.7 Elastic vs. thermal deployment of stent
-p
re
One could argue as the reasons to why using the thermal shape memory (SME) property of Nitinol is

a better approach than using its superelastic property. In order to answer this question, we ran another
lP

set of simulations and compared the results obtained when using each of the properties.
na

Taking advantage of Nitinol (super)elasticity allows the stent to be easily compressed into the thin
ur

catheter and expand by many folds to deploy into the diseased aortic region. In this study, the stent

expanded from an 8 mm to a 30 mm diameter at 301 K temperature, less than the 310K body temper-
Jo

ature.

If the shape memory effect property is used, a room temperature (≈21o C or lower) is sufficient to com-

press the stent and keep it compressed. During deployment, the stent should be kept cold (i.e. below

the transition temperature) and it will self-expand as it becomes in contact with the body (≈36.8o C).

Theoretically, stents can be manufactured to react (transition) at temperatures slightly above of 310

K body temperature (say at 313 K, for example). But these stents would need to be heated to expand

which in practice, is difficult to accomplish. This is one of the major limitation for thermal deploy-

17
Journal Pre-proof

ment.

The stress contours (refer Supplementary Figures, Fig. S4) show that when Nitinol superelastic prop-

erty is used, the stent is subjected to 0.325 MPa, i.e. higher stress than when the thermal property is

used (0.3 MPa as seen in Fig. 2). The results also show that regions of higher stress are generated at

the interface artery-stent when the elastic property is used. In the thermal case, we showed that after

the stent is released from the catheter, residual strains can be seen, which will eventually vanish as the

of
stent becomes in contact with the body temperature (transition temperature 310 K).

ro
-p
The difference in stress recorded may be due to the kind of expansion process. In superelastic de-
re
ployment (we mean when the superelastic property of Nitinol is used), the stent expands and seals

the surface when the loads are removed whereas the stent will regain its original shape in two steps
lP

during thermal deployment. Once the stent is released from the catheter (i.e. when the mechanical
na

load is removed), the stent still have some residual strains which will not result in sealing the surface

yet. The complete sealing will occur during the second step, during which the stent own temperature
ur

increases due to being exposed to the body temperature. At that point, the stent expands to its original
Jo

size.

This higher stress on the stent recorded during superelastic deployment may damage the aortic walls.

This finding has been supported by experimental data. In an experimental study conducted by Kurib-

ayashi et al. [43], on Nitinol stents, the team reported that during the superelastic deployment, the

stent expansion was so quick that the video recording could not catch the deployment. They also con-

ducted experiments using SME property of Nitinol which showed that the stent deployment was more

gentle. This is in agreement with our simulations.

18
Journal Pre-proof

5 Conclusions

In this work, we designed a closed-cell honeycomb stent-graft structure such as that used in endovas-

cular aortic (or aneurysm) repair (EVAR) to treat abdominal aortic aneurysm (AAA). We employed

the Abaqus/Simulia FEM commercial software to evaluate the strains on the structure to sufficiently

compress and fit it into a thin catheter. We also evaluated the radial displacements and the von Mises

of
stress once the stent fully expanded. We chose Nitinol for the stent material and Dacron fabric for

ro
the graft. The influence of geometric parameters such as the strut spacing and the strut radius of

curvature on the crimping strains was studied. The effect of the blood flow on the stent-graft was
-p
investigated using the FSI built-in feature in Abaqus/Simulia. Deployment and expansion of the stent
re
were simulated using the shape-memory effect (SME) of Nitinol and associated sealing stresses were
lP

recorded.
na

This study supported the following conclusions:


ur

• the crimping strains increased linearly as the stent strut spacing increased. Likewise, for the strut
Jo

radius of curvature

• the sealing stresses exerted by the Nitinol stent on the aortic walls were below the materials yield

stress, far away enough to not be concerned with stent fracture

• the wall shear stresses developed on the stented aorta were close to those obtained of an healthy

aorta. This indicated that the stent was adhering well to the aortic walls and that migration was not

likely to occur

• using Nitinol thermo-reactivity as opposed to its superelasticity was useful, allowing the stent-

graft to gently deploy, perhaps owing to the gradual temperature increase in contrast to the sudden

19
Journal Pre-proof

removal of loads when using its superelasticity property

• future work will involve simulations of the stent in a patient-specific anatomy of the aorta with

aneurysm.

Acknowledgment

of
This publication was made possible by the NPRP award # 7-032-2-016 from the Qatar National Re-

ro
search Fund (a member of Qatar Foundation). The statements made herein are solely the responsibility

of the authors.
-p
re
lP

References
na

[1] H. Banno, F. Cochennec, J. Marzelle, and J. P. Becquemin. Comparison of fenestrated endovas-

cular aneurysm repair and chimney graft techniques for pararenal aortic aneurysm. J. Vasc.
ur

Surg., 60:31–39, 2014.


Jo

[2] S. W. K. Cheng. Novel endovascular procedures and new developments in aortic surgery. British

Journal of Anaesthesia, 117:ii3–ii2, 2016.

[3] J. W. V. Keulen, K. L. Vincken, J. V. Prehn, J. L. Tolenaar, L. W. Bartels, M. A. Viergever, F. L.

Moll, and J. A. V. Herwaarden. The Influence of Different Types of Stent Grafts on Aneurysm

Neck Dynamics after Endovascular Aneurysm Repair. Euro. J. vasc. Endo.surg, 39:193–199,

2010.

[4] D. Daye and T. G. Walker. Complications of endovascular aneurysm repair of the thoracic and

abdominal aorta:evaluation and management. Cardiovasc Diagn Ther., 8:S138–S156, 2018.

20
Journal Pre-proof

[5] R. O. Tadros, P. L. Faries, S. H. Ellozy, R. A. Lookstein, A. G. Vouyouka, R. Schrier, J. Kim,

and M. L. Marin. The impact of stent graft evolution on the results of endovascular abdominal

aortic aneurysm repair. J. Vasc. Surg., 59:1518–27, 2014.

[6] P. McKavanagh, G. Zawadowski, N. Ahmed, and M. Kutryk. The evolution of coronary stents.

Expert Rev. Cardiovasc. Ther, 16:219–228, 2018.

[7] J. M. Jani, M. Leary, A. Subic, and M. A. Gibson. A review of shape memory alloy research,

of
applications and opportunities. Mat. Design, 56:1078–1113, 2014.

ro
[8] M. J. Mahtabi, N. Shamsaei, and M. R. Mitchell. Fatigue of Nitinol: The state-of-the-art and

ongoing challenges. J. Mech. Behavior Bio. Mat., 50:228–254, 2015.


-p
[9] S. Mahdis and C. Youngjae. An overview of thin film nitinol endovascular devices. Acta Bio-
re
materialia, 21:20–34, 2015.
lP

[10] D. J. McGrath, B. O’Brien, M. Bruzzi, N. Kelly, J. Clauser, U. Steinseifer, and P. E. McHugh.

Evaluation of cover effects on bare stent mechanical response. J. Mech. Behavior Bio. Mat.,
na

61:567–580, 2016.
ur

[11] PET. https://www.britannica.com/science/


Jo

polyethylene-terephthalate.

[12] G. P. Kumar and F. Cui. Stent design parameters and crimpability. Int. J. Cardiology, 223:552–

553, 2016.

[13] A. Schiavone, T. Y. Qiu, and L. G. Zhao. Crimping and deployment of metallic and polymeric

stents – finite element modelling. Vessel Plus, 1:12–21, 2017.

[14] M. F. Urbano, A. Cadelli, F. Sczerzenie, P. Luccarelli, S. Beretta, and A. Coda. Inclusions Size-

based Fatigue Life Prediction Model of NiTi Alloy for Biomedical Applications. Shape Memory

and Superelasticity, 1(2):240–251, Jun 2015.

21
Journal Pre-proof

[15] A. Gerbay, J. Terreaux, A. Cerisier, M. Vola, and K. Isaaz. Impact of very high pressure stent

deployment on angiographic and long-term clinical outcomes in true coronary bifurcation le-

sions treated by the mini-crush stent technique: A single center experience. Int. Heart Journal,

69:32–36, 2017.

[16] R. Blic, F. Cochennec, F. Alomran, H. Kobeiter, E. Allaire, P. Desgranges, and J. P. Becquemin.

Impact of Stent-Graft Oversizing on Gutter Areas after Chimney Graft Repair for Complex

of
Abdominal Aortic Aneurysms. Annals of Vasc. Surg., 51:200–206, 2018.

ro
[17] M. S. Kim and L. S. Dean. In-stent restenosis. Cardiovasc. Ther., 29:190–198, 2011.

[18] L. H. Timmins, M. W. Miller, F. J. Clubb, and J. E. Moore. Increased artery wall stress post-
-p
stenting leads to greater intimal thickening. Lab Invest., 91:955–967, 2011.
re
[19] F. R. Arko, E. H. Murphy, C. M. Davis, E. D. Johnson, S. T. Smith, and C. K. Zarins. Dynamic
lP

geometry and wall thickness of the aortic neck of abdominal aortic aneurysms with intravascular

ultrasonography. J. Vasc. Surg., 46:891–897, 2007.


na

[20] M. Amabili, P. Balasubramanian, I. Breslavsky, G. Ferrari, and E. Tubaldi. Viscoelastic charac-


ur

terization of woven dacron for aortic grafts by using direction-dependent quasi-linear viscoelas-
Jo

ticity. J. Mech. Behavior Bio. Mat., 82:282–290, 2018.

[21] G. Ferrari, P. Balasubramanian, E. Tubaldi, F. Giovanniello, and M. Amabili. Experiments on

Dynamic Behaviour of a Dacron Aortic Graft in a Mock Circulatory Loop. J. Biomech., 2019.

https://doi.org/10.1016/j.jbiomech.2019.01.053.

[22] S. Zhao, L. Gu, and S. R. Froemming. Finite Element analysis of the Implantation of a Self-

Expanding Stent: Impact of Lesion Calcification. J. Med. Dev., 6:021001–1, 2012.

[23] D. Martin and F. Boyle. Finite element analysis of balloon-expandable coronary stent deploy-

ment: influence of angioplasty balloon configuration. Int. J. Numer. Methods.Biomed. Eng.,

22
Journal Pre-proof

29:1161–1175, 2013.

[24] C. Kleinstreuer, Z. Li, C. A. Basciano, S. Seelecke, and M. A. Farber. Computational mechanics

of nitinol stent grafts. J. Biomechanics, 41:2370–2378, 2008.

[25] R. Jayendiran, B. M. Nour, and A. Ruimi. Computational fluid-structure interaction analysis of

blood flow on patient-specific reconstructed aortic anatomy and aneurysm treatment with dacron

graft. J. Fluids. Struct., 81:693–711, 2018.

of
[26] Y. C. Fung. Biomechanics:Mechanical Properties of Living Tissues. Springer-Verlag P. 568

ro
ISBN 0-387-97947-6, 1993.

[27] Shell elements. https://www.linkedin.com/pulse/


-p
shells-vs-solids-finite-element-analysis-quick-review-kuusisto-p-e-/.
re
[28] R. Jayendiran, B. M. Nour, and A. Ruimi. Fluid-structure interaction (FSI) analysis of stent-
lP

graft for aortic endovascular aneurysm repair (EVAR): Material and structural considerations. J.

Mech. Behavior Bio. Mat., 87:95–110, 2018.


na

[29] D. Stoeckel, C. Bonsignore, and S. Duda. A survey of stent designs. Min. Inv. Thera. Allied
ur

Tech., 11:137–147, 2002.


Jo

[30] T. C. Gasser, R. W. Ogden, and G. A. Holzapfel. Hyperelastic modelling of arterial layers with

distributed collagen fibre orientations. J. R. Soc. Interface, 3:15–35, 2006.

[31] G. A. Holzapfel. Determination of material models for arterial walls from uniaxial extension

tests and histological structure. Journal of Theoretical Biology, 238:290–302, 2006.

[32] R. Jayendiran, B. M. Nour, and A. Ruimi. Dacron graft as replacement to dissected aorta: A

three-dimensional fluid-structure- interaction analysis. J. Mech. Behavior Bio. Mat., 78:329–

341, 2018.

[33] C. A. Bustos, C. M. G. Herrera, and D. J. Celentano. Mechanical characterisation of Dacron

23
Journal Pre-proof

graft: Experiments and numerical simulation. J. Biomechanics, 49:13–18, 2016.

[34] E. Tubaldi, M. Amabili, and M. P. Paidoussis. Fluid-structure interaction for nonlinear response

of shells conveying pulsatile flow. Journal of Sound and Vibration, 371:252–376, 2016.

[35] M. A. Jankowska, M. Bartkowiak-Jowsa, and R. Bedzinski. Experimental and constitutive mod-

eling approaches for a study of biomechanical properties of human coronary arteries. Journal

of the Mechanical Behaviour of Biomedical Materials, 50:1–12, 2015.

of
[36] T. J. Pedley. The fluid mechanics of large blood vessels. Cambridge University Press, Cam-

ro
bridge, 1980.

[37] K. Perktold, M. Resch, and H. Florian. Pulsatile non-newtonian flow characteristics in a three-
-p
dimensional human carotid bifurcation model. J. Biomech. Eng., 113:464–475, 1991.
re
[38] E. Tubaldi, M. P. Paidoussis, and M. Amabili. Nonlinear dynamics of Dacron aortic prostheses
lP

conveying pulsatile flow. J. Biomech. Eng., 240:061004, 2018.

[39] R. Tuzi and P. Blondeaux. Intermittent turbulence in a pulsating flow. J. Fluids. Mech., 599:51–
na

79, 2008.
ur

[40] D. Lagoudas, D. Hartl, Y. Chemisky, L. Machado, and P. Popov. Constitutive model for the
Jo

numerical analysis of phase transformation in polycrystalline shape memory alloys. Int. J. Plas-

ticity, 32:155–183, 2012.

[41] M. A. Qidwai and D. C. Lagoudas. Numerical implementation of a shape memory alloy ther-

momechanical constitutive model using return mapping algorithms. Int. J. Numer. Meth Engng.,

47:1123–1168, 2000.

[42] Abaqus version 6.13. Analysis User’s Manual. Dassault Systems Simulia Corp, 2013.

[43] K. Kuribayashi, K. Tsuchiya, Z. You, D. Tomus, M. Umemoto, T. Ito, and M. Sasaki. Self-

deployable origami stent grafts as a biomedical application of Ni-rich TiNi shape memory alloy

24
Journal Pre-proof

foil. Mater. Sci. Engg. A, 419:131–137, 2006.

[44] M. S. Cabrera, C. W. J. Oomens, and F. P. T. Baaijens. Understanding the requirements of

self-expandable stents for heart valve replacement: Radial force, hoop force and equilibrium. J.

Mech. Behavior Bio. Mat., 68:252–264, 2017.

[45] D. C. Lagoudas. Shape Memory Alloys. Springer, 2008.

[46] T. Pham, M. Deherrera, and W. Sun. Finite element analysis of the biomechanical interaction

of
between coronary sinus and proximal anchoring stent in coronary sinus annuloplasty. Comput.

ro
Methods Biomech. Biomed. Eng., 17:1617–1629, 2014.

[47] E. S. Di Martino, G. Guadangi, A. Fumero, G. Ballerini, R. Spirito, P. Boglioli, and A. Redaelli.


-p
Fluid-structure interaction within realistic three-dimensional models of the aneurysmatic aorta
re
as a guidance to assess the risk of rupture of the aneurysm. Med. Eng. Phy, 23:647–655, 2001.
lP

[48] F. M. Callaghan and S. M. Grieve. Normal patterns of thoracic aortic wall shear stress

measured using four-dimensional flow MRI in a large population. Am. J. Phys., 2018.
na

https://doi.org/10.1152/ajpheart.00017.2018.
ur

[49] S. R. Bailey. DES Design: Theoretical Advantages and Disadvantages of Stent Strut Materials,
Jo

Design, Thickness, and Surface Characteristics. J. Interven. Cardiol, 22:S3–S17, 2009.

[50] H. zahedmanesh and C. Lally. Determination of the influence of stent strut thickness using the

finite element method: implications for vascular injury and in-stent restenosis. Med. Bio. Eng.

Comp, 2009. https://doi.org/10.1007/s11517-009-0432-5.

25
Journal Pre-proof

FIGURE CAPTIONS

Fig.1 Nitinol stent maximum crimping strain as a function of strut curvature and strut spacing

Fig.2 Residual strains after unloading (left); von Misses stresses (in Pa) on stented aorta with aneurysm

(right) during thermal deployment

Fig.3 Normal contact force between the aortic walls and the stent during the stent deployment for

various stent diameters

of
Fig.4 Input pressure profile used in the simulations (top). Radial displacement (in m), von Mises stress

ro
(in Pa) and WSS (in Pa) contours (bottom) on stented aorta as a result of blood flow. Values are

recorded at systolic pressure


-p
re
Supplementary Figures
lP

Fig.S1 Fluid-structure interaction cosimulation scheme (a) and Gauss-Seidel time marching algorithm
na

(b) used in Abaqus/Simulia

Fig.S2 Model validation: cross section (a) stent model (b) crimping strain (c) of diamond cell stent
ur

Fig.S3 Nitinol phase transformation diagram [45]


Jo

Fig.S4 Von Misses stresses (in Pa) on stented aorta with aneurysm at contact and when completely

sealed during superelastic deployment

26
Journal Pre-proof

H=12mm H=6mm
H=12mm

6.5mm 6.5mm 6.5mm


0.5mm 0.5mm 0.5mm

Rc=1.5mm Rc=1.5mm Rc=1mm

of
Y Y
ro Y
Z
-p Z Z
H=12 mm, Rc =1.5 mm H=6 mm; Rc =1.5 mm H=12 mm; Rc =1.0 mm
re
10 10
Crimping Strain (%)
Crimping strain (%)

lP

9.5 9.5

9 9

8.5
na

8.5

8
8 1 1.2 1.4 1.6
4 6 8 10 12
Strut spacing, H (mm) Radius of curvature, Rc (mm)
ur

Fig. 1. Nitinol stent maximum crimping strain as a function of strut curvature and strut spacing
Jo

27
Journal Pre-proof

+3.046e+05
+1.834e+05
+5.632e+04
+6.143e+03 Y
Y +9.146e+02
Z +7.574e+01 Z
Unloaded +2.932e+00 Initial contact between Completely
Configuration +0.000e+00 the stent and the aorta wall sealed

of
Fig. 2. Residual strains after unloading (left); von Misses stresses (in Pa) on stented aorta with aneurysm (right)
during thermal deployment

ro
-p
re
lP
na
ur
Jo

28
Journal Pre-proof

15

Contact force (N)


10

0
25 30 35 40
Stent diameter (mm)

Fig. 3. Normal contact force between the aortic walls and the stent during the stent deployment for various stent
diameters

of
ro
-p
re
lP
na
ur
Jo

29
Journal Pre-proof

110 Di Martino et al.,2001

Pressure, p (mmHg)
Present Model

100

90

80

70
0 0.5 1 1.5
Time, t (s)
Pressure profile

of
ro
+1.448e-03 +2.046e+08
-p +4.210e+00
+8.812e-04 +6.428e+07 +4.200e+00
+6.531e-04 +6.632e+05 +4.186e+00
re
+4.643e-04 +6.143e+05 +4.168e+00
+5.345e-05 +9.146e+04 Y +4.145e+00
+2.730e-05
Y +7.574e+03 +4.121e+00 Y
+6.896e-06 +2.932e+02 +4.113e+00
+2.782e-06 Z X +1.481e+01 Z X +4.108e+00 Z X
lP

Radial displacement Von Mises stress Wall Shear Stress


Fig. 4. Input pressure profile used in the simulations (top). Radial displacement (in m), von Mises stress (in Pa)
and WSS (in Pa) contours (bottom) on stented aorta as a result of blood flow. Values are recorded at systolic
na

pressure
ur
Jo

30
Journal Pre-proof

Supplementary Figures

of
ro
-p
re
lP
na
ur
Jo

31
Journal Pre-proof

CFD FEA
Pressure Stress
Shear Cosimulation Strain
Engine
Mesh (CSE) Displacement
Morphing Velocity

(a)
i i+1 i+2
CFD
pi ui pi+1 ui+1

FEA

of
i i+1 i+2
(b)
Fig. S1 Fluid-structure interaction cosimulation scheme (a) and Gauss-Seidel time marching

ro
algorithm (b) used in Abaqus/Simulia
-p
re
lP
na
ur
Jo

32
Journal Pre-proof

0.35mm
L
4mm
9.5mm

+9.018e-02
+8.336e-02
+7.431e-02

of
+6.243e-02
0.28mm +4.929e-02
Y +2.814e-02
+9.912e-03
Y

ro
Z X +4.905e-04 Z
(a) (b) (c)
Fig. S2 Model validation: cross section (a) stent model (b) crimping strain (c) of diamond cell stent
-p
re
lP
na
ur
Jo

33
Journal Pre-proof

260 K 273 K 301 K 316 K

A
Mf As

Ms Af

Austenite
Fig. S3 Nitinol phase transformation diagram [45]

of
ro
-p
re
lP
na
ur
Jo

34
Journal Pre-proof

+3.256e+05
+2.032e+05
+6.124e+04
+8.012e+03 Y
+9.316e+02
+7.875e+01 Z
+3.512e+00 Initial contact between Completely
+0.000e+00 the stent and the aorta wall sealed
Fig. S4 Von Misses stresses (in Pa) on stented aorta with aneurysm at contact and when completely

of
sealed during superelastic deployment

ro
-p
re
lP
na
ur
Jo

35
Journal Pre-proof

Highlights

 Crimping strain increased as strut spacing and radius of curvature increased.


 Sealing stresses on the aortic walls well below the yield stress of aortic walls.
 WSS developed on the stented aorta were close to those obtained of an healthy
aorta.
 Contact forces shows stent was adhering well to the aortic walls and no migration
 Nitinol SME allows the stent-graft to gently deploy smoothly compared to superelastic

of
ro
-p
re
lP
na
ur
Jo
Figure 1
Figure 2
Figure 3
Figure 4

You might also like