Deployment of a Bulk Metallic Glass-Based Self ... - ACS Publications

Sep 14, 2016 - In the following subsections, we describe the generation of the patient-specific descending aorta model, generation of the stent model,...
0 downloads 0 Views 1MB Size
Subscriber access provided by Northern Illinois University

Article

Deployment of a bulk metallic glass based selfexpandable stent in a patient-specific descending aorta Gideon Praveen Kumar, Mehdi Jafary Zadeh, and Fangsen Cui ACS Biomater. Sci. Eng., Just Accepted Manuscript • DOI: 10.1021/acsbiomaterials.6b00342 • Publication Date (Web): 14 Sep 2016 Downloaded from http://pubs.acs.org on September 20, 2016

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

ACS Biomaterials Science & Engineering is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 24

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Biomaterials Science & Engineering

Deployment of a bulk metallic glass based selfexpandable stent in a patient-specific descending aorta Gideon Praveen Kumar1, Mehdi Jafary-Zadeh1* and Fangsen Cui1 1

Institute of High Performance Computing, A*STAR (1 Fusionopolis Way #16-16 Connexis

Singapore 138632) E-mail: *[email protected]

KEYWORDS: self-expandable stent; bulk metallic glass; finite element analysis ; crimping; deployment; aorta; nitinol ABSTRACT: The emergence of bulk metallic glasses (BMGs) has been tantalizing in biomedical applications such as development of novel cardiovascular stents. Numerous investigations have confirmed the superior functional properties and biocompatibility of BMGs over conventional crystalline alloys as stent materials. However, a detailed understanding of the mechanical behavior of BMG-based stents during different stages of their application is still scarce. Here, by quantitative finite element analyses (FEA), we explore the deployment process of a BMG-based self-expandable stent in a patient specific descending aorta to evaluate the arterial stresses and the vessel deformation during the stent deployment. We further benchmark the performance of the BMG based stent by comparing the deployment results both with the results of a similar nitinol based stent and the experimental failure strength of the human arteries. Our detailed analyses confirm that the

1 ACS Paragon Plus Environment

ACS Biomaterials Science & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 24

proposed BMG stent can be safely deployed in the artery without vessel over-stretching and mechanical failure, preventing unexpected vessel injuries and resultant pathological responses. Our findings would be insightful for further investigations towards realization of novel BMG-based stent applications. INTRODUCTION: As a result of continuously increasing aging population and the need for improvement of living standards, there is a prevalent demand for development of novel biomaterials to enhance the longevity and quality of human life. In the quest for novel biomaterials, both their functional and mechanical properties must be carefully evaluated to guarantee long-term biocompatibility and structural integrity of the medical device. The former (biocompatibility) addresses an important caveat which is the highly corrosive environment in the human body for many conventional materials and alloys, leading to release of toxic ions of some constitutive metallic elements1-3. Hence, biocompatibility is an essential requirement for biomaterials to avoid adverse and toxic effects on the human system. The latter (structural integrity), which poses certain requirements on the mechanical properties of biomaterials, addresses concerns regarding the complex physiological loading by surrounding tissues, blood vessel walls and bones during the operational life of any implantable medical device especially during different disease conditions. Monolithic bulk metallic glasses (BMGs), also known as “liquid-like” metals, have an amorphous microstructure, i.e. lack of atomistic long-range order4, so that they do not contain microscopic structural defects like grain boundaries and dislocations. In the absence of defect driven mechanisms of plasticity, BMGs offer extraordinary mechanical properties such as large elastic strain limit, high strength and resilience (i.e. high elastic energy storage), improved wear resistance, fatigue endurance, and excellent high temperature formability, in comparison to their crystalline counterparts. Besides, studies reporting biocompatibility of BMGs have been on the rise1,

5-8

. Another feature of BMGs which makes it suitable for

2 ACS Paragon Plus Environment

Page 3 of 24

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Biomaterials Science & Engineering

medical device applications is their high imaging compatibility9-10. For example it has been shown that BMGs have better MRI compatibility compared to 316L stainless steel. Owing to the combination of such superior and desirable structural and functional properties, BMGs are in the cutting-edge of materials research for biotechnological applications, including stent devices.11,1, 10, 12-14. Stents are widely employed to restore the patency of narrowed or clogged blood vessel and other human conduits and as device frames in prosthetic heart valve replacement. Despite the success of conventional polycrystalline biometals as stent material, the emergence of BMGs has opened new horizons for development of competitive alternatives for different stent applications9 . The functional properties and biocompatibility of a prototypical Zr-based BMG to be used as stent material have been extensively investigated, and it is confirmed that this class of materials offers better biocompatibility than conventional crystalline alloys in both vitro cell culture and in vivo animal implantations.4, 7, 14-16 The high corrosion resistance of Zr-based BMGs is due to the passive films primarily composed of ZrO2 that are formed on the alloy’s outer surface.10,

16

On the other hand, the high strength of the BMGs greatly

enables the production of stents with thinner struts5, which contributes to its ease of deliverability and reduces the rate of restenosis. For example, in comparison to traditional 316L stainless steel stents, Zr-based BMG stents would require just one-third of the strut cross-section and yet provide in excess of five times the strength.10

However, in contrast

with numerous studies on the aforementioned functional properties of BMGs for stent applications, a complete understanding of the mechanical behavior of BMG-based stents during different stages of their delivery and service is still scarce. A recent study has reported results of finite element based fatigue analysis of a BMG-based vascular stent subjected to pulsatile fatigue loading and showed that the stent can withstand pulsatile pressure loads from blood and the vessel wall9. For the clinical application of BMGs as self-

3 ACS Paragon Plus Environment

ACS Biomaterials Science & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 24

expandable stent materials, mechanical assessment of the stent is crucial to assure the integrity of the device during the crimping procedure. Therefore, we have recently conducted in-depth analyses on the crimpability of BMG-based stents for several percutaneous applications17 requiring self-expandable stents. Our analyses demonstrated that BMG-based self-expandable stents containing diamond shaped cells are susceptible to catastrophic failure through shear banding during the crimping stage17. We further deduced that BMGs could be used in non-diamond shaped stents like the Zenith stent graft which is a woven stent where shear banding was not observed during crimping. Figure S1 in the Supplementary Material provides a comparison of strains between our simulated nitinol- and BMG-based Zenith stents in the crimped profile, indicating that the integrity of both stents are preserved during the crimping procedure. Though crimpability is very important, where the stent is progressively compressed to a smaller diameter without failure, its real time deployment in a patient specific blood vessel is equally important which determines the final outcome of the actual surgical procedure. Alarmingly high arterial stresses and vessel stretching, post stent deployment could result in profound long-term histological response including exuberant neointimal proliferation and luminal stenosis18. Hence the aim of this work mainly is to simulate the deployment of a BMG-based Zenith TX2 TAA19 stent in a patient specific descending aorta to quantitatively evaluate the arterial stresses and the vessel deformation due to stent deployment. Since nitinol is currently the most popular self-expandable stent material20-25, we have benchmarked the performance of our BMG stent with that of nitinol. Here, it is noteworthy that computational modeling and simulation (CM&S) has been widely used as a research and development tool to support medical device applications by analysing the integrity, safety and effectiveness of the medical devices. Currently, variety of computational techniques is available such as boundary element method (BEM), finite volume method (FVM), finite difference method (FDM) and finite element method (FEM).

4 ACS Paragon Plus Environment

Page 5 of 24

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Biomaterials Science & Engineering

As the specific application of these techniques, BEM is predominantly used for sound propagation analysis and electromagnetics26-27, FVM is used in computational fluid dynamics28, and FDM is used for certain case-based specific problems29-30. Besides, FEM has proved to be a formidable tool that can be used to analyze ubiquitous biomechanical problems including stent applications and explore the effects of design parameters on the mechanical performance of stents and their resulted stress distributions on the artery wall31-34. Hence, owing to its solid theoretical foundation and computational efficiency, most of the stent related theoretical work available in the literature has used FEM to conduct stent structural analyses which have also been experimentally confirmed in a variety of cases22, 3541

. Our FE analyses are also based on the available documentations provided by The Food

and Drug Administration (FDA or USFDA) on stent analysis42.

MATERIALS & METHODS: We used ABAQUS (v. 6.13-2) (Dassault Systemes, MA) to conduct the finite element analysis and we focused on the stent deployment where the following parts were modeled in the simulations. •

a patient-specific descending aorta;



stent;



crimper.

In the following subsections, we describe the generation of the patient specific descending aorta model, generation of the stent model, nitinol, BMG and artery material modeling and FEA settings.

Geometry models and meshes The Zenith TX2 TAA Endovascular stent Graft19, 43-45 (Cook Medical Europe, Bjaeverskov, Denmark) as shown in Fig. 1 (a) was studied. This cylindrical graft comprises of a metallic

5 ACS Paragon Plus Environment

ACS Biomaterials Science & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 24

stent sutured to a polyester fabric with braided polyester and monofilament polypropylene suture. Since the focal point of this work is the metallic stent, the fabric was not considered in this work. CAD model of the bare metal stent used in this study is shown in Fig. 1(b). It is a woven stent with an outer diameter of 30 mm and a wire diameter is 0.3 mm. The stent was meshed with 6636 linear shear flexible B31 beam elements (Fig. 1(c)), which account for large axial strains and transverse shear strains32-33. Typically, woven stents (e.g. the BMGbased Zenith stent used in this study) are modeled using circular beam elements19, 32-33 as it has been shown that the computation time of beam element is reduced to one third of that of continuum element, where both formulations exhibit comparable results46-47. Moreover, the usage of beam elements in woven stenting simulations has also been successfully validated experimentally48. We processed the DICOM CT images of an aortic model (ascending, arch and the descending) where the geometrical information of the patient specific aorta was extracted as axial images and then imported to Mimics (Materialise, Leuven, Belgium). Segmentation, removal of secondary branches and smoothing of surfaces were done using Mimics before exporting it out as STL file. After some in-house geometric editing and clean-up the STL was then imported into ABAQUS as shown in Fig. 1 (d). 5118 quadrilateral shell elements [(general purpose with reduced integration (S4R)31] were used for modeling the artery since these elements tend to resemble the geometric characteristics of the artery wall wherein the artery dimension in the thickness direction is much smaller than that along the other directions. Moreover, the inherent plane stress assumption for shell elements helps in simplification of the equations and aids in convergence of the solution for this highly nonlinear problem, especially when combining the complex tissue material model with the nonlinear patient specific three dimensional (3D) aorta anatomy49.

Shell elements are

increasingly being used for complicated patient specific artery models to reduce the

6 ACS Paragon Plus Environment

Page 7 of 24

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Biomaterials Science & Engineering

computational time31, 49-50. A Boolean operation was then done to remove the ascending aorta and the arch of aorta to get the patient-specific descending aorta. Since the medical images contain no information on the vessel wall thickness, we assigned the thickness of shell elements to be 2 mm51-53. Fig. 1(e) & (f) shows the final patient-specific descending aorta model and its mesh respectively. The outer diameters of the proximal and the distal ends of the descending aorta were 18 mm and 16 mm respectively. This change in the diameter is because the lumen of the aorta tapers in a way such that the cross-sectional area at the beginning of the descending aorta becomes almost 50% of the area at the aortic root54. This can be expressed by the relationship

‫ܣ‬ሺܼ ሻ = ‫ܣ‬଴ ݁ ି௞௭/௥ (1)

where k is the taper factor, r is the vessel radius, z is the axial distance of the vessel from the aortic root, and A0 is the area at the aortic root. The tapering phenomenon observed in the aorta can be evidently seen in Fig. 1(d)

Material models Nitinol Model

The nonlinear super elasticity of nitinol was modeled as a thermo-mechanical coupled superelastic plastic model implemented by an inbuilt user-defined UMAT for ABAQUS/Standard finite element solver. The model is based on an additive strain decomposition55, in which the total strain is the sum of the elastic strain, the transformation strain, and the plastic strain as follows ߝ = ߝா௅ + ߝ்ோ + ߝ௉௅ (2)

7 ACS Paragon Plus Environment

ACS Biomaterials Science & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 24

Even though plastic strains develop as soon as the material is loaded beyond full transformation, exposing it to temperatures above it austenite finish [Af] temperature causes a thermal transformation which makes it recover the strain55-56. The transformation strain is around 6 to 8%, but the elastic strain is much smaller, ranging from 2 to 4 %. Since the transformation strains are large compared to typical elastic strains in nitinol, the material is said to be super elastic57-58. Hence unlike conventional alloys like stainless steel, in nitinol, recovery of up to 10% strains is possible59-61.

The UMAT used in this work is written with several user-defined material constants that define the stress and temperature induced phase transformations occurring between the austenite and martensite phases. Initially, nitinol exhibits austenite crystalline structure with a Young’s modulus of 40000 MPa and Poisson’s ratio of 0.33. When loaded beyond the start of transformation load which typically occurs during stent crimping, nitinol transforms from austenite to martensite. This austenite to twinned martensite transformation is driven by the resolution of shear forces, and happens within a range of stress levels that are characteristic of the material35. The transformation strain can be expresses as,

ߝ்ோ = ߙ∆ߞሺߜ‫ܨ‬/ߜߪሻ, ‫ ܨ‬ௌ ≤ ‫ ܨ ≤ ܨ‬ி

(3)

where ζ is the fraction of the martensite phase, α is a material constant which can be calibrated from uniaxial tensile tests, and F is a transformation potential. The trend is similar for the reverse transformation but at different stress levels. The intensity of the transformation is based on a stress potential law given by ∆ߞ = ݂ሺߪ, ߞሻΔ‫ܨ‬

(4)

8 ACS Paragon Plus Environment

Page 9 of 24

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Biomaterials Science & Engineering

For the manufacturer specific nitinol20 we used in this analysis, the transformation is initiated from a transformation start-load of 449 MPa and ends at a transformation-end load of 540 MPa into a complete martensite phase with a Young’s modulus of 32000 MPa and Poisson’s ratio of 0.33. Martensite begins to transform back to austenite once unloading begins at 250 MPa which happens when the crimped stent is let to self-expand in a blood vessel where the body temperature is typically above the Af. The stent recovers to its original shape with full Austenite reversal which occurs at a transformation unloading stress of 83 MPa. BMG Model

The BMG was modelled as a linear elastic material with Young’s modulus of 80 GPa and Poisson’s ratio of 0.38 obtained from experimental results9. Indeed, using a linear elastic model to represent BMGs is reasonable since they are typically brittle and behave almost as linear elastic materials exhibiting no global plasticity during deformation up to their failure6263

. Furthermore, this material’s elastic strain, yield strength, and resilience (elastic energy

storage per unit volume) is much higher than many other conventional metals and alloys. Especially, the high resilience of BMGs plays an important role in their elastic recovery due to their high effective elastic strain which results in their spring back phenomenon64-65. Artery Model

The mechanical behavior of the descending aorta was modeled using a homogenous, isotropic hyperelastic constitutive model. The constitutive law is based on a reduced polynomial strain energy density function of third order. The 3rd order polynomial form of the strain energy density function U can be written as

ܷ = ‫ܥ‬ଵ଴ ሺ‫ܫ‬ଵ − 3ሻ + ‫ܥ‬ଶ଴ ሺ‫ܫ‬ଵ − 3ሻଶ + ‫ܥ‬ଷ଴ ሺ‫ܫ‬ଵ − 3ሻଷ

(5)

9 ACS Paragon Plus Environment

ACS Biomaterials Science & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 24

where I1 is the first invariant of the Cauchy – Green tensor

‫ܫ‬ଵ = ߣଵଶ + ߣଶଶ + ߣଶଷ , ߣ௜ = ‫ܬ‬ଵ/ଷ ߣ௜

(6)

where λi are the principal stretches and J is the total volume ratio. The corresponding material coefficients33 for the polynomial hyperelastic model used in Eq. 5 are C10 = 0.0104673 MPa, C20 = 0.0194098 MPa and C30 = 0.0109830 MPa, which were obtained from experimental results of a realistic aorta which is the blood vessel studied in this work33.

Simulation strategy for stent deployment

The crimper was modeled as a rigid deformable body with 10100 4-mode surface elements (SFM3D4). Axial and circumferential boundary conditions were used to constrain the translational (Uz) and rotational (Uθ) degrees of freedom of the stent and artery models to allow deformation only in the radial direction. Mesh sensitivity studies were performed to ensure that all results were independent of further mesh refinements (For details, see Figure S2 in Supplementary Material and the discussions therein). The numerical analysis is nonlinear and involves large deformations which were considered while defining the step in the ABAQUS setup. Our simulation consists of two steps. •

Stent crimping - here the 30 mm stent was progressively crimped to 8 mm resembling being inserted into a 24F sheath by reducing the diameter of the crimper in a controlled manner. In this step, contact was modeled between the stent and inner surface of the crimper.

10 ACS Paragon Plus Environment

Page 11 of 24

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Biomaterials Science & Engineering



Stent deployment - from the crimped configuration, the stent was allowed to reexpand by progressively letting the crimper recoil with the stent against the vessel wall. Here, an additional contact is modeled between the stent and the inner surface of the descending aorta.

We used a surface-to-surface algorithm in order to model the interactions between the different parts and used a frictionless contact between crimper inner surface and the stent and a friction of 0.0566 between the stent and the descending aorta. Fig. 2 shows the different stages of stent deployment in the patient-specific descending aorta. RESULTS & DISCUSSION: Fig. 3 (a) and (b) show the contour plots for the distribution of stresses generated in the arterial walls of descending aorta upon stent deployment made of nitinol and BMG, respectively. It is observed that in both cases there is a base level of stress generated by the overall expansion of the artery. Meanwhile, there are higher stresses of 0.26 MPa and 0.33 MPa due to nitinol and BMG stent deployment respectively, near the contact interfaces where there is stretching of the tapered aorta caused by the expansion of the more uniform circular stent. This is intuitive since there is a large radial deformation against the artery wall. This is because, after crimping, typically the stent is allowed to “self-expand” progressively inside the artery driven by its own stored strain energy. It expands until it reaches an equilibrium state as a function of the vessel’s elasticity. More specifically, upon stent-vessel contact, the vessel begins to expand due to the stent’s outward force till the stent and the vessel reach a stress equilibrium while the stent is in compression and the vessel is in tension25. Once a stress equilibrium state is reached, typically, the stent takes the shape of the vessel as its final form, thereby, getting embedded in the vessel musculature and securing good anchorage and vessel apposition. Fig. 3(c) shows the deformed stent which has taken the shape of the descending aorta which is similar for both nitinol and BMG-based stents. This phenomenon of the stent getting embedded into the vessel musculature is called “stent

11 ACS Paragon Plus Environment

ACS Biomaterials Science & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 24

penetration”67. In other words, penetration is the peak radial displacement of the artery on its inner surface due to contact between the stent and the artery after stent deployment. Once the stress equilibrium is reached, penetration can be calculated as the difference between the artery initial inner diameter and the deformed stent outer diameter. Our results showed that penetration induced by the nitinol and BMG stents were 0.32 mm and 0.36 mm, respectively. This difference in penetration can be attributed to the fact that though both nitinol and BMG stents have the same initial (crimped) geometries, their intrinsic mechanical properties in terms of elasticity and deployment mechanisms are different. Nitinol has a unique elastic hysteresis as a result of its martensite to austenite transformation that allows the stent to reexpand upon deployment onto the vessel wall 35, 68-70 . In this case, the stent pushes outwards and positions itself against the artery wall with a low chronic outward force [COF]71. On the other hand, in the case of the BMG stent, the penetration is a result of intrinsic pure elasticity (with no phase transformation) of the stent behaving in analogy with the push-back behavior of a spring released from the compressed state. Here, we recall that BMGs are exceptionally high resilient materials in general, which substantially allows them to store higher amount of elastic energy per unit volume than any other 1500 conventional engineering metals and alloys.65, 72 This implies that to obtain a certain amount of penetration upon deployment, a smaller stents (and hence less material) is needed if the stent is made of metallic glasses rather than other conventional alloys. Extensive work has been done on the deployment of self-expandable33, 68-70 stents in human arteries. However, based on the best of our knowledge, there is a lack of information about direct comparison of the axial and circumferential stresses with experimental results. Researches mostly have just reported the maximum stresses observed in the artery upon stent deployment. In this work, we compare our simulated axial and circumferential stresses in the patient-specific descending aorta post stent implantation directly with existing experimental

12 ACS Paragon Plus Environment

Page 13 of 24

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Biomaterials Science & Engineering

data73 for the ultimate strength of human artery73. This comparison is crucial since during and after stent deployment there could be over stretching and mechanical failure of the arteries which in turn could lead to vessel injuries as well as profound long-term histological response including exuberant neointimal proliferation18. Hence, keeping the vessel stresses below its ultimate strength is imperative73-75. The histogram depicted in Fig. 3(d), compares the axial and circumferential stresses in the aorta due to the deployment of modeled nitinol and BMGbased stents with the ultimate failure strength of the human artery. This figure clearly demonstrates that both the axial and circumferential stresses due to the BMG stent deployment are comparable with those of nitinol stent deployment, and are much lower than the failure limit of human arteries. Hence, it can be concluded that upon careful design, the BMG-based stents can to be used safely during the crimping and deployment procedures without their own mechanical failure and without causing any injury to the blood vessel, implying that BMGs are promising candidates for future stent applications. It is noteworthy that while the material models used in this work to simulate BMG9, nitinol76 and the human artery33 are all obtained from experimental results that are already available in the literature, further experimental studies to evaluate our FE data in real life clinical applications, is our ongoing research as a progression of the current work.

SUMMARY: The objective of this work was to check the feasibility of using BMGs as stent material for self-expandable stent applications. Stent deployment, which is an important process that decides the outcome of the surgical process, was considered here. Finite element modeling and simulation of woven stent deployment in a patient-specific descending aorta were presented. A prototypical Zr-based BMG was used as the stent material and benchmarked against nitinol since currently, nitinol is the most popular material for selfexpandable stent applications. Nonlinear FEA was used to determine the stresses in the

13 ACS Paragon Plus Environment

ACS Biomaterials Science & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 24

vessel during and after stent deployment. The obtained simulation results were compared with experimental data for ultimate failure strength of artery vessels. The effect of stent material on the arterial stresses and the stent penetration in the artery were demonstrated. We found that arterial stresses and penetration as a result of a BMG stent deployment was comparable with those of nitinol stent. Additionally, the induced stresses were lower than the failure limit of the arteries which is an encouraging finding of this study. Further to our previous details of crimping analysis, our simulation results of deployment indicate that BMGs can be safely used as a stent material for developing woven stents as they can be crimped and deployed without much difficulty which is very important for the success of any percutaneous procedure. As the progression of the current work, ongoing research on design optimization of BMG-based woven stents is proposed in order to exploit the mechanical properties of BMGs and obtain better performance than nitinol based stents. Our work lays the platform to come up with new BMG based stent designs that could be used for various percutaneous applications requiring self-expanding stents.

ACKNOWLEDGEMENTS: The authors gratefully acknowledge the financial support from the Agency for Science, Technology and Research (A*STAR), Singapore and the use of computing resources at the A*STAR Computational Resource Centre, Singapore.

14 ACS Paragon Plus Environment

Page 15 of 24

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Biomaterials Science & Engineering

References

1. Espallargas, N.; Aune, R.; Torres, C.; Papageorgiou, N.; Muñoz, A., Bulk metallic glasses (BMG) for biomedical applications—A tribocorrosion investigation of Zr 55 Cu 30 Ni 5 Al 10 in simulated body fluid. Wear 2013, 301 (1), 271-279. 2. Ortiz, A. J.; Fernández, E.; Vicente, A.; Calvo, J. L.; Ortiz, C., Metallic ions released from stainless steel, nickel-free, and titanium orthodontic alloys: toxicity and DNA damage. American Journal of Orthodontics and Dentofacial Orthopedics 2011, 140 (3), e115-e122. 3. Köster, R.; Vieluf, D.; Kiehn, M.; Sommerauer, M.; Kähler, J.; Baldus, S.; Meinertz, T.; Hamm, C. W., Nickel and molybdenum contact allergies in patients with coronary in-stent restenosis. The Lancet 2000, 356 (9245), 1895-1897. 4. Zhu, S.; Xie, G.; Qin, F.; Wang, X.; Inoue, A., Ni-and Be-free Zr-based bulk metallic glasses with high glass-forming ability and unusual plasticity. Journal of the mechanical behavior of biomedical materials 2012, 13, 166-173. 5. Horton, J.; Parsell, D. In Biomedical Potential of a Zirconium-Based Bulk Metallic Glass, MRS Proceedings, Cambridge Univ Press: 2002; p CC1. 5. 6. He, W.; Chuang, A.; Cao, Z.; Liaw, P. K., Biocompatibility study of zirconium-based bulk metallic glasses for orthopedic applications. Metallurgical and Materials Transactions A 2010, 41 (7), 1726-1734. 7. Hua, N.; Huang, L.; Wang, J.; Cao, Y.; He, W.; Pang, S.; Zhang, T., Corrosion behavior and in vitro biocompatibility of Zr–Al–Co–Ag bulk metallic glasses: An experimental case study. Journal of Non-Crystalline Solids 2012, 358 (12), 1599-1604. 8. Liu, Y.; Wang, Y.-M.; Pang, H.-F.; Zhao, Q.; Liu, L., A Ni-free ZrCuFeAlAg bulk metallic glass with potential for biomedical applications. Acta biomaterialia 2013, 9 (6), 7043-7053. 9. Huang, L.; Pu, C.; Fisher, R. K.; Mountain, D. J.; Gao, Y.; Liaw, P. K.; Zhang, W.; He, W., A Zr-based bulk metallic glass for future stent applications: Materials properties, finite element modeling, and in vitro human vascular cell response. Acta biomaterialia 2015, 25, 356-368. 10. Li, H.; Zheng, Y., Recent advances in bulk metallic glasses for biomedical applications. Acta biomaterialia 2016, 36, 1-20. 11. Temenoff, J. S.; Mikos, A. G., Biomaterials: the intersection of biology and materials science. 2014. 12. Liu, L.; Qiu, C.; Chen, Q.; Chan, K.; Zhang, S., Deformation behavior, corrosion resistance, and cytotoxicity of Ni‐free Zr‐based bulk metallic glasses. Journal of Biomedical Materials Research Part A 2008, 86 (1), 160-169. 13. Zhu, S.; Wang, X.; Qin, F.; Inoue, A., A new Ti-based bulk glassy alloy with potential for biomedical application. Materials Science and Engineering: A 2007, 459 (1), 233-237. 14. Gu, X.; Zheng, Y.; Zhong, S.; Xi, T.; Wang, J.; Wang, W., Corrosion of, and cellular responses to Mg–Zn–Ca bulk metallic glasses. Biomaterials 2010, 31 (6), 1093-1103. 15. Zhu, S.; Wang, X.; Xie, G.; Qin, F.; Yoshimura, M.; Inoue, A., Formation of Ti-based bulk glassy alloy/hydroxyapatite composite. Scripta Materialia 2008, 58 (4), 287-290. 16. Huang, L.; Zhu, C.; Muntele, C. I.; Zhang, T.; Liaw, P. K.; He, W., Surface engineering of a Zr-based bulk metallic glass with low energy Ar-or Ca-ion implantation. Materials Science and Engineering: C 2015, 47, 248-255. 17. Praveen Kumar, G.; Jafary-Zadeh, M.; Tavakoli, R.; Cui, F., Feasibility of using bulk metallic glass for self-expandable stent applications. Journal of Biomedical Materials Research Part B: Applied Biomaterials 2016, n/a-n/a. DOI: 10.1002/jbm.b.33718.

15 ACS Paragon Plus Environment

ACS Biomaterials Science & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 24

18. Zhao, H. Q.; Nikanorov, A.; Virmani, R.; Jones, R.; Pacheco, E.; Schwartz, L. B., Late stent expansion and neointimal proliferation of oversized nitinol stents in peripheral arteries. Cardiovascular and interventional radiology 2009, 32 (4), 720-726. 19. Demanget, N.; Avril, S.; Badel, P.; Orgéas, L.; Geindreau, C.; Albertini, J.-N.; Favre, J.-P., Computational comparison of the bending behavior of aortic stent-grafts. Journal of the mechanical behavior of biomedical materials 2012, 5 (1), 272-282. 20. Kumar, G. P.; Cui, F.; Phang, H. Q.; Su, B.; Leo, H. L.; Hon, J. K. F., Design considerations and quantitative assessment for the development of percutaneous mitral valve stent. Medical engineering & physics 2014, 36 (7), 882-888. 21. Kumar, G. P.; Cui, F.; Danpinid, A.; Su, B.; Hon, J. K. F.; Leo, H. L., Design and finite element-based fatigue prediction of a new self-expandable percutaneous mitral valve stent. Computer-Aided Design 2013, 45 (10), 1153-1158. 22. Kleinstreuer, C.; Li, Z.; Basciano, C.; Seelecke, S.; Farber, M., Computational mechanics of Nitinol stent grafts. Journal of biomechanics 2008, 41 (11), 2370-2378. 23. Duerig, T.; Pelton, A.; Stöckel, D., An overview of nitinol medical applications. Materials Science and Engineering: A 1999, 273, 149-160. 24. Duerig, T.; Tolomeo, D.; Wholey, M., An overview of superelastic stent design. Minimally invasive therapy & allied technologies 2000, 9 (3-4), 235-246. 25. Duerig, T.; Wholey, M., A comparison of balloon-and self-expanding stents. Minimally invasive therapy & allied technologies 2002, 11 (4), 173-178. 26. Prakash, R.; Education, S., Analysis Of Electromagnetic Pulse Simulators. 2009. 27. Brebbia, C. A.; Popov, V., Boundary Elements and Other Mesh Reduction Methods XXXIII. Wit Press: 2011. 28. Moukalled, F.; Mangani, L.; Darwish, M., The Finite Volume Method in Computational Fluid Dynamics: An Advanced Introduction with OpenFOAM® and Matlab. Springer: 2015; Vol. 113. 29. Dolicanin, C.; Nikolic, V.; Dolicanin, D., Application of finite difference method to study of the phenomenon in the theory of thin plates. Appl. Math. Inform. And Mech 2010, 2, 1. 30. Prieto, F. U.; Muñoz, J. J. B.; Corvinos, L. G., Application of the generalized finite difference method to solve the advection–diffusion equation. Journal of Computational and Applied Mathematics 2011, 235 (7), 1849-1855. 31. Arokiaraj, M. C.; Palacios, I. F., Finite element modeling of a novel self-expanding endovascular stent method in treatment of aortic aneurysms. Scientific reports 2014, 4. 32. Zhao, S.; Gu, L.; Froemming, S. R., Finite element analysis of the implantation of a self-expanding stent: impact of lesion calcification. Journal of Medical Devices 2012, 6 (2), 021001. 33. Zhao, S.; Liu, X. C.; Gu, L., The impact of wire stent fabrication technique on the performance of stent placement. Journal of Medical Devices 2012, 6 (1), 011007. 34. Auricchio, F.; Conti, M.; De Beule, M.; De Santis, G.; Verhegghe, B., Carotid artery stenting simulation: from patient-specific images to finite element analysis. Medical engineering & physics 2011, 33 (3), 281-289. 35. Azaouzi, M.; Makradi, A.; Belouettar, S., Deployment of a self-expanding stent inside an artery: a finite element analysis. Materials & Design 2012, 41, 410-420. 36. McGrath, D.; Bruzzi, M.; McHugh, P., Nitinol stent design–understanding axial buckling. Journal of the mechanical behavior of biomedical materials 2014, 40, 252-263. 37. Wu, W.; Yang, D.-Z.; Qi, M.; Wang, W.-Q., An FEA method to study flexibility of expanded coronary stents. Journal of Materials Processing Technology 2007, 184 (1), 447450.

16 ACS Paragon Plus Environment

Page 17 of 24

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Biomaterials Science & Engineering

38. Bobel, A.; Petisco, S.; Sarasua, J. R.; Wang, W.; McHugh, P., Computational bench testing to evaluate the short-term mechanical performance of a polymeric stent. Cardiovascular Engineering and Technology 2015, 1-14. 39. Kumar¹, G.; Leo, H.; Cui, F., Effect of strut curvature on the crimpability of mitral valve stents. 40. Kumar, G. P.; Kabinejadian, F.; Liu, J.; Ho, P.; Leo, H. L.; Cui, F., Simulated Bench Testing to Evaluate the Mechanical Performance of New Carotid Stents. Artificial organs 2016. 41. Ismail, M.; Kumar, G. P.; Kabinejadian, F.; Nguyen, Y. N.; Cui, F.; Tay, E. L. W.; Leo, H. L., An Experimental and Computational Study on the Effect of Caval Valved Stent Oversizing. Cardiovascular Engineering and Technology 2016, 1-16. 42. Food; Administration, D., Guidance for Industry and FDA Staff, Non-Clinical Engineering Tests and Recommended Labeling for Intravascular Stents and Associated Delivery Systems. Rockville, MD: US 2013. 43. Hsu, H.-L.; Chen, C.-K.; Chen, P.-L.; Chen, I.-M.; Hsu, C.-P.; Chen, C.-W.; Shih, C.C., The impact of bird-beak configuration on aortic remodeling of distal arch pathology after thoracic endovascular aortic repair with the Zenith Pro-Form TX2 thoracic endograft. Journal of vascular surgery 2014, 59 (1), 80-88. 44. Perrin, D.; Demanget, N.; Badel, P.; Avril, S.; Orgéas, L.; Geindreau, C.; Albertini, J. N., Deployment of stent grafts in curved aneurysmal arteries: toward a predictive numerical tool. International journal for numerical methods in biomedical engineering 2015, 31 (1). 45. Weidman, J. M.; Desai, M.; Iftekhar, A.; Boyle, K.; Greengard, J. S.; Fisher, L. M.; Thomas, R. L.; Zannetti, S., Engineering Goals for Future Thoracic Endografts—How Can We Make Them More Effective? Progress in cardiovascular diseases 2013, 56 (1), 92-102. 46. Ballew, W.; Seelecke, S. In A comparison of FE beam and continuum elements for typical nitinol stent geometries, SPIE Smart Structures and Materials+ Nondestructive Evaluation and Health Monitoring, International Society for Optics and Photonics: 2009; pp 72890R-72890R-11. 47. Mortier, P.; De Beule, M.; Van Loo, D.; Masschaele, B.; Verdonck, P.; Verhegghe, B., Automated generation of a finite element stent model. Medical & biological engineering & computing 2008, 46 (11), 1169-1173. 48. Kim, J. H.; Kang, T. J.; Yu, W.-R., Mechanical modeling of self-expandable stent fabricated using braiding technology. Journal of biomechanics 2008, 41 (15), 3202-3212. 49. Zhang, Y.; Dunn, M. L.; Hunter, K. S.; Lanning, C.; Ivy, D. D.; Claussen, L.; Chen, S. J.; Shandas, R., Application of a microstructural constitutive model of the pulmonary artery to patient-specific studies: validation and effect of orthotropy. Journal of biomechanical engineering 2007, 129 (2), 193-201. 50. Gayzik, F. S.; Bostrom, O.; Duma, S. M.; Stitzel, J. D., AN EXPERIMENTAL AND FINITE ELMENT STUDY OF THE PORCINE CAROTID ARTERY UNDER DYNAMIC LOADING. 51. Li, A. E.; Kamel, I.; Rando, F.; Anderson, M.; Kumbasar, B.; Lima, J. A.; Bluemke, D. A., Using MRI to assess aortic wall thickness in the multiethnic study of atherosclerosis: distribution by race, sex, and age. American Journal of Roentgenology 2004, 182 (3), 593597. 52. Mensel, B.; Kühn, J.-P.; Schneider, T.; Quadrat, A.; Hegenscheid, K., Mean thoracic aortic wall thickness determination by cine MRI with steady-state free precession: validation with dark blood imaging. Academic radiology 2013, 20 (8), 1004-1008. 53. Erbel, R.; Eggebrecht, H., Aortic dimensions and the risk of dissection. Heart 2006, 92 (1), 137-142.

17 ACS Paragon Plus Environment

ACS Biomaterials Science & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 24

54. Leondes, C. T., Biomechanical Systems: Techniques and Applications, Volume II: Cardiovascular Techniques. CRC press: 2000. 55. Zadno, G.; Duerig, T. In Linear and non-linear superelasticity in NiTi, Proceedings of the MRS International Meeting on Advanced Materials., 1988; pp 201-206. 56. Stoeckel, D., The Shape Memory Effect: Phenomenon, Alloys and Applications. Nitinol Devices and Components, Inc. Línea, Consultado en 2012, 06-13. 57. Rebelo, N.; Walker, N.; Foadian, H. In Simulation of implantable nitinol stents, Abaqus user’s conference, 2001; pp 1-14. 58. Lecce, L., Shape Memory Alloy Engineering: For Aerospace, Structural and Biomedical Applications. Elsevier: 2014. 59. Robertson, S.; Pelton, A.; Ritchie, R., Mechanical fatigue and fracture of Nitinol. International Materials Reviews 2012, 57 (1), 1-37. 60. Stoeckel, D.; Pelton, A.; Duerig, T., Self-expanding nitinol stents: material and design considerations. European radiology 2004, 14 (2), 292-301. 61. Pelton, A., Nitinol fatigue: a review of microstructures and mechanisms. Journal of Materials Engineering and Performance 2011, 20 (4-5), 613-617. 62. Huang, Y.; Khong, J.; Connolley, T.; Mi, J., The onset of plasticity of a Zr-based bulk metallic glass. International Journal of Plasticity 2014, 60, 87-100. 63. Aydıner, C. C.; Üstündag, E., Residual stresses in a bulk metallic glass cylinder induced by thermal tempering. Mechanics of Materials 2005, 37 (1), 201-212. 64. Lee, M.; Park, E.; Ott, R.; Kim, B.; Eckert, J., Evaluation of the relationship between the effective strain and the springback behavior during the deformation of metallic glass ribbons. Applied Physics Letters 2014, 105 (6), 061906. 65. Tian, L.; Cheng, Y.-Q.; Shan, Z.-W.; Li, J.; Wang, C.-C.; Han, X.-D.; Sun, J.; Ma, E., Approaching the ideal elastic limit of metallic glasses. Nature communications 2012, 3, 609. 66. Ragkousis, G. E.; Curzen, N.; Bressloff, N. W., Simulation of longitudinal stent deformation in a patient-specific coronary artery. Medical engineering & physics 2014, 36 (4), 467-476. 67. Borghi, A.; Murphy, O.; Bahmanyar, R.; McLeod, C., Effect of stent radial force on stress pattern after deployment: a finite element study. Journal of Materials Engineering and Performance 2014, 23 (7), 2599-2605. 68. Wu, W.; Qi, M.; Liu, X.-P.; Yang, D.-Z.; Wang, W.-Q., Delivery and release of nitinol stent in carotid artery and their interactions: a finite element analysis. Journal of biomechanics 2007, 40 (13), 3034-3040. 69. Wu, W.; Wang, W.-Q.; Yang, D.-Z.; Qi, M., Stent expansion in curved vessel and their interactions: a finite element analysis. Journal of biomechanics 2007, 40 (11), 25802585. 70. Azaouzi, M.; Lebaal, N.; Makradi, A.; Belouettar, S., Optimization based simulation of self-expanding Nitinol stent. Materials & Design 2013, 50, 917-928. 71. Morgan, N., Medical shape memory alloy applications—the market and its products. Materials Science and Engineering: A 2004, 378 (1), 16-23. 72. Greer, A. L., Metallic glasses… on the threshold. Materials Today 2009, 12 (1), 1422. 73. Teng, Z.; Tang, D.; Zheng, J.; Woodard, P. K.; Hoffman, A. H., An experimental study on the ultimate strength of the adventitia and media of human atherosclerotic carotid arteries in circumferential and axial directions. Journal of biomechanics 2009, 42 (15), 25352539. 74. Lally, C.; Dolan, F.; Prendergast, P., Cardiovascular stent design and vessel stresses: a finite element analysis. Journal of biomechanics 2005, 38 (8), 1574-1581.

18 ACS Paragon Plus Environment

Page 19 of 24

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Biomaterials Science & Engineering

75. Fortier, A.; Gullapalli, V.; Mirshams, R. A., Review of biomechanical studies of arteries and their effect on stent performance. IJC Heart & Vessels 2014, 4, 12-18. 76. Gong, X.-Y.; Pelton, A. R. In Finite element analysis on nitinol medical applications, ASME 2002 International Mechanical Engineering Congress and Exposition, American Society of Mechanical Engineers: 2002; pp 439-440.

19 ACS Paragon Plus Environment

ACS Biomaterials Science & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 24

Figure 1 CAD geometries and meshes used in this study. (a) Zenith TX2 TAA Endovascular Graft (www.cookmedical.com); (b) Isometric view of the CAD model of the bare metal wire based stent; (c) Stent meshed with beam elements; (d) CAD model of the patient-specific aorta; (e) CAD model of the region of interest, i.e. the patient-specific descending aorta; (f) Patient-specific descending aorta meshed with quadrilateral shell elements.

20 ACS Paragon Plus Environment

Page 21 of 24

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Biomaterials Science & Engineering

Figure 2 Steps involved in the simulation of the deployment process (a) The crimper, stent and the artery are assembled in a cylindrical co-ordinate system; (b) The crimper is used to crimp the stent, where its diameter is reduced from 30 mm to 8 mm; (c) The stent is let to expand into the vessel till a stress equilibrium is achieved between the stent and the vessel.

21 ACS Paragon Plus Environment

ACS Biomaterials Science & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 22 of 24

22 ACS Paragon Plus Environment

Page 23 of 24

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Biomaterials Science & Engineering

Figure 3 Vessel and stent configuration after deployment (a) Contour plot of von Mises stress distribution (S, von Mises) in the post deployment descending aorta when the stent material is nitinol; (b) Contour plot of von Mises stress distribution (S, von Mises) in the post deployment descending aorta when the stent material is BMG; (c) The deformed stent which has taken the shape of the vessel after full deployment; (d) Histogram comparing the axial and circumferential stresses in the artery after stent deployment with nitinol and BMG as the stent material against experimental test results on the failure strength on human arteries.

23 ACS Paragon Plus Environment

ACS Biomaterials Science & Engineering

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 24 of 24

For Table of Contents Use Only Deployment of a bulk metallic glass based self-expandable stent in a patientspecific descending aorta Gideon Praveen Kumar1, Mehdi Jafary-Zadeh1* and Fangsen Cui1 1

Institute of High Performance Computing, A*STAR, Singapore 138632

24 ACS Paragon Plus Environment