Sei sulla pagina 1di 15

Biomaterials 103 (2016) 278e292

Contents lists available at ScienceDirect

Biomaterials
journal homepage: www.elsevier.com/locate/biomaterials

Review

Micro and nanotechnologies in heart valve tissue engineering


Anwarul Hasan a, b, c, **, John Saliba b, Hassan Pezeshgi Modarres d, e, f, Ahmed Bakhaty f,
Amir Nasajpour c, Mohammad R.K. Mofrad f, g, Amir Sanati-Nezhad d, e, *
a

Department of Mechanical and Industrial Engineering, College of Engineering, Qatar University, Doha 2713, Qatar
Department of Mechanical Engineering, Faculty of Engineering and Architecture, American University of Beirut, Beirut 1107 2020, Lebanon
c
Biomaterials Innovation Research Center, Division of Biomedical Engineering, Department of Medicine, Brigham and Women's Hospital, Harvard Medical
School, Cambridge, MA 02139, USA
d
BioMEMS and Bioinspired Microuidic Laboratory, Department of Mechanical and Manufacturing Engineering, University of Calgary, Calgary, Canada
e
Center for BioEngineering Research and Education, University of Calgary, Calgary, Canada
f
Molecular Cell Biomechanics Laboratory, Departments of Bioengineering and Mechanical Engineering, University of California Berkeley, 208A Stanley Hall,
Berkeley, CA 94720-1762, USA
g
Physical Biosciences Division, Lawrence Berkeley National Lab, Berkeley, CA 94720, USA
b

a r t i c l e i n f o

a b s t r a c t

Article history:
Received 4 April 2016
Received in revised form
26 June 2016
Accepted 1 July 2016
Available online 2 July 2016

Due to the increased morbidity and mortality resulting from heart valve diseases, there is a growing
demand for off-the-shelf implantable tissue engineered heart valves (TEHVs). Despite the signicant
progress in recent years in improving the design and performance of TEHV constructs, viable and
functional human implantable TEHV constructs have remained elusive. The recent advances in micro and
nanoscale technologies including the microfabrication, nano-microber based scaffolds preparation, 3D
cell encapsulated hydrogels preparation, microuidic, micro-bioreactors, nano-microscale biosensors as
well as the computational methods and models for simulation of biological tissues have increased the
potential for realizing viable, functional and implantable TEHV constructs. In this review, we aim to
present an overview of the importance and recent advances in micro and nano-scale technologies for the
development of TEHV constructs.
2016 Elsevier Ltd. All rights reserved.

Keywords:
Heart valves
Tissue engineering
Nanotechnologies
Microtechnologies
Scaffolds
Hydrogels

1. Introduction
Nearly 300,000 valve replacement surgeries are performed
every year due to heart valve diseases [1e3]. Replacement surgeries
are often required due to severe valve damages that render valve
repair not feasible. Replacements valves can be mechanical or
bioprosthetic, each having its set of severe limitations. Mechanical
valves may result in thrombosis or blood clotting and hence, they
require the patient to be on anti-blood clotting drugs for the rest of
their life. Bioprosthetic valves, on the other hand, lack longevity and
are known to calcify, which leads to stiffening and thickening of the

* Corresponding author. BioMEMS and Bioinspired Microuidic Laboratory,


Department of Mechanical and Manufacturing Engineering, University of Calgary,
2500 University Drive NW, Calgary, Alberta T2N 1N4, Canada.
** Corresponding author. Department of Mechanical and Industrial Engineering,
College of Engineering, Qatar University, Doha 2713, Qatar.
E-mail addresses: ahasan@qu.edu.qa, mh211@aub.edu.lb (A. Hasan), amir.
sanatinezhad@ucalgary.ca (A. Sanati-Nezhad).
http://dx.doi.org/10.1016/j.biomaterials.2016.07.001
0142-9612/ 2016 Elsevier Ltd. All rights reserved.

valve cusps and eventually insufcient closure, resulting in leakages. Additionally, these valves neither allow for somatic growth
nor remodeling after implantation. The lack of remodeling proves
to be a major problem for children as they would require repeated
replacements every few years to accommodate physical growth [4].
Tissue engineering promises to overcome these challenges and to
provide suitable replacement valves helping both growth and selfremodeling capabilities in vivo [5e10].
Various approaches have been developed to synthesize TEHVs
heart valves for clinical uses, including: 1) the use of regenerative
capability of the body (in-situ) to recruit circulating cells into
implanted decellularized or preshaped biodegradable synthetic
scaffolds, 2) hybrid approaches via modication of decellularized
scaffolds using biomaterials and hydrogels to improve cell seeding
onto scaffold, and 3) development of the valve substitutes in vitro
using prolonged tissue culture prior to implantation [11,12]. The
method of generating implantable TEHVs starts with seeding
appropriate cells on biodegradable scaffolds and culturing them
under appropriate environmental cues in bioreactors, controlling

A. Hasan et al. / Biomaterials 103 (2016) 278e292

the arrangement, directionality, and polarity of cells regulating the


cell-matrix interaction to induce tissue formation, and nally
implanting the preconditioned constructs in vivo. The goal is to
develop TEHVs that demonstrate biophysical, biochemical, and
biomimetic remodeling properties like those of native valves [13].
Micro and nanoscale technologies have emerged to signicantly
contribute in not only incorporating helpful nano-microscale features in novel biomaterials but also in engineering biomimetic
cellular microenvironment in the tissue preconditioning bioreactors. Examples include, 1) ensuring microscale porous structure
and nanoscale dimensional accuracies in engineered biomaterials;
2) controlling the topography for the accurate delivery of growth
factors or genes to specic cells in a certain region of tissue; 3)
synthesis of microstructures with anisotropic arrangement and
layer by layer fabrication supporting multi-dimensional biomimetic anisotropic mechanical properties [14e16]; micromolding
for the fabrication of complex geometries of TEHVs [17,18]; and 5)
controlled delivery of chemicals, biomolecules, and non-viral gene
carriers through incorporating microuidics [19].
In this review, we present applications of the micro and nanoscale technologies to the development of TEHV constructs with the
focus on the pulmonary and aortic heart valves. We further discuss
the challenges for translation of engineered heart valves to the
clinic and future directions for the development of next generation
TEHVs.

2. Macro and microscale structure of native heart valves


The four heart valves (pulmonary, tricuspid, mitral, and aortic
valves) lie on the valvular basal plan, as demonstrated in Fig. 1(a).
The aortic valve is responsible for controlling blood ow from the
left ventricle to the aorta. The pulmonary valve regulates ow from
the right ventricle to the pulmonary artery, and the mitral and
tricuspid valves regulate blood ow to the left and right atria,
respectively.

279

The aortic and pulmonary valves are composed of three semilunar leaets (or cusps) along with sinus complexes called valve
roots as shown in Fig. 1(b) and (c). These valve leaets are
composed of four main components, namely the commissural region, the belly of the leaets, the lannula with the noduli of Arantii,
and the coapting surface [20]. The leaets are thin, exible and trilayered structures that are composed of collagen, elastin, and glycosaminoglycans (GAGs). The three sublayers are the brosa, the
spongiosa, and the ventricularis [21], as shown in Fig. 1(d). The
intricate structure of the valves results in strong mechanical
properties necessary to withstand the high trans-valvular pressures, and the low exural stiffness required for regular valve
operation. The annulus of the heart valves consists of a dense
elastin and collagen meshwork. The collagenous bers in the
corrugated brosa layer are shown in Fig. 1(e) [20].
The thickest layer, the brosa, consists mainly of a dense
network of corrugated type-I collagen bers arranged in the
circumferential direction. The highly organized collagen of the
brosa provides the primary load-bearing properties of the heart
valve [20]. The collagen bers form sheets in the brosa which
allows expansion of the leaet during closure of the heart valve
[24]. The middle layer, the spongiosa, is composed of highly hydrated GAGs and proteoglycans (PGs) as well as loosely arranged
collagen and elastin. The spongiosa acts as a cushion between the
brosa and the ventricularis and enables shearing between the two
layers during loading and unloading, and absorb the load by
transferring it to the elastic aortic wall, resulting in minimal stress
on the leaet itself. In addition, the tissue is nearly incompressible
due to the physical properties of bound water molecules that
comprise almost 70% of the natural weight. It is believed that the
viscoelastic and mechanical-force damping properties of the tissue
is due to water retained by GAGs [25].
Below the inow surface is the thinnest layer, the ventricularis,
which is comprised of a dense network of collagen bers and
elastin sheets, Fig. 1(d) and (e). The collagen in ventricularis is less

Fig. 1. Heart valve structure and composition: a) structure of heart valves, the schematics of the 2D position of the four valves on valvular basal plane of heart where P: pulmonary
valve, AO: aortic valve, M: mitral valve, and T: tricuspid valve, bee) Composition of the heart valves: b) the schematic cross section of an aortic valve leaet, c) histology of the cross
section of a valve leaet showing the three main layers, brosa, spongiosa and ventricularis, d) The schematic of the elastin and collagen microstructures in different layers during
systolic and diastolic cycles, e) Arrangement of collagen bers as well as distribution of elastin and GAG's. Figures adapted and reprinted from Ref. [6,22,23], with permissions from
the Royal Society of Chemistry and Elsevier Science.

280

A. Hasan et al. / Biomaterials 103 (2016) 278e292

organized compared to that in brosa and contain radially oriented


elastin bers [26]. The ventricularis helps to reduce large radial
strains during the high blood ow over the valves when they are
fully opened. The observed nonlinear stress-strain response of
heart valve tissue is a consequence of its microstructure. Under
small strains, little force is required to stretch the crimped collagen
and elastin bers. During larger strains, the bers become
uncrimped and require a much larger force (i.e., they are much
stiffer). The ventricularis and the brosa are tightly connected
enabling the heart valve to function optimally [26].
Two types of cells mainly make up the heart valve tissue:
valvular interstitial cells (VICs) and valvular endothelial cells
(VECs). Some nerve cells exist to varying degrees but are very
sparse and their functions are not well understood [9]. VICs are
phenotypically of broblast cell types [27] however, they often
exhibit the characteristics of myobroblasts or some other cell
types. VECs cover the leaet surfaces, aligned orthogonal to the
blood ow and are responsible for the production of GAGs.
3. Tissue engineering of heart valves
Tissue engineering aims to solve the issues of organ failure by
replacing diseased tissues or organs with biomimetic replicates as a
whole or part [28]. Cell-laden scaffold replicates are expected to
produce functional living valves with reasonable compositional,
architectural, and mechanical similarity to native analogues. Three
essential elements needed for the synthesis of a functional TEHVs
are scaffolds from natural or synthetic biomaterials, cells encapsulated in scaffolds, and signaling molecules regulating cells
response [29]. The incorporated cells might be stem cells, progenitor cells or differentiated cells such as VECs or VICs. Cells are expected to secrete the extracellular matrix (ECM), therefore
replacing the biodegradable polymer or decellularized scaffold
with natural ECM of the heart valves or reinforces the polymer
structure [30]. The cells can be harvested from the individual patients' tissue or recruited from bloodstream to ensure the
maximum adaptability between the TEHV and the patient's body
[29].
For a functional scaffold for TEHVs, in addition to general requirements of biomaterials like biocompatibility, biodegradability,
and being non-immunogenic, some additional features are needed
to be considered [31]. The TEHVs should be non-obstructive, nonthrombogenic, and resistant to calcication, with the ability to
grow, remodel, and adapt to the changing physiological conditions,
and complete and prompt opening and closure in patients' body
[32e34]. In addition, they need to be adaptable to periodic changes
in dimensions, shape, and stress during opening/closure of the
cyclical beating and hence tensile-shear-exural loading [35]. The
valves need to hold robust biomechanical properties to withstand
the hemodynamic pressures and ows of the cardiac environment.
TEHVs should have the ability to respond to biological signals and
initiating self-repair [36]. Furthermore, to avoid thickening of the
leaet the degradation of the scaffold materials must occur in
parallel to the ECM deposition. Despite an extensive research in
classical tissue engineering of hart valves and promising results in
animal models, the TEHVs have not been introduced safely and
reliably to human patients. The reasons for this include the
complexity of the procedure, the sub-optimal performance in vivo,
the valve tissue retraction [37], and the inter-patient variability
[38].
In-situ tissue regeneration provides biophysical and biochemical
cues to recruit cells gradually from the bloodstream into decellularized or synthetic scaffolds and to produce valve tissue at the site
of destination [33,38e44]. The main advantages of this method are
simplied fabrication protocols, and usage of the intrinsic

regenerative nature of the living body to induce repopulation of the


host cells and tissue remodeling. The in-situ approaches also do not
need an extensive in vitro cell production in bioreactors [45,46].
However, in-situ TEHVs still suffer from the immunogenic risk
factors, shortage of biocompatible grafts, and the lack of controlled
cell engraftment to deposit load-bearing ECM and develop anisotropic tissue properties [47]. The efciency and performance of the
in-situ approach strongly depend on the host body, particularly,
when the patient suffers from immunologic defects or metabolic
diseases, whereas the calcication could be serious [6,39,48]. The
drawbacks of in-situ models necessitate the employment of smart
biomaterials to regulate the behavior of recruited cells and to
provide the formation of a functional heart valve tissues durable
under biomechanical loading [49]. Biodegradable materials have
been utilized for guided tissue regeneration of heart valves
(Fig. 2(a)) [41], although they are inversely inuenced by technological challenges. Novel hybrid polymeric heart valve scaffolds
merge the advantages of anatomically shaped decellularized scaffolds and benets of synthetic scaffolds, providing a physiological
function for hybrid heart valves (Fig. 2(c)) [50,51]. For instance,
hybrid heart valve scaffolds were synthesized via combining bioresorbable polymers like poly(4-hydroxybu-tyrate) and poly(3hydroxybutyrate-co4-hydroxybutyrate)
with
decellularized
porcine aortic heart valve scaffold to enhance the biomechanical
behavior of tissue function [52]. Hybrid heart valve tissues were
also developed by electrospinning-based coating of broblast
growth factors and chitosan and mesenchymal stem cells onto
porcine aortic heart valve leaets. The developed hybrid strategy
increased signicantly cell recruitment, biomimetic ECM production, and the strength of heart valves [53]. Luminal grafts seeded
with autologous cells and coated with chemoattractants like
vascular endothelial growth factor (VEGF), or stromal cell-derived
factor-1 (SDF-1a) have shown promising results to eliminate
immunogenicity, thrombus formation, and calcication [54].
Despite all advances in the development of in-situ and hybrid
TEHVs, there is slight progress for the development of biomimetic
TEHVs that can withstand the pressure inside the heart for an
extended period of time. Most of the TEHVs, nevertheless, require
advanced bioactive sites to be able to organize the interaction between cells and scaffolds. Endothelial cells were not able to adhere
rmly to the scaffold and lost under both blood ow and mechanical loading, followed by activation of the inammatory
response, wall thickening, size remodeling, and thrombosis
[24,52,55,56]. Instead, in vitro TEHV models provide exibility for
engineering cells composition, controlling biophysical and
biochemical properties of polymeric biomaterial, and regulating
the interaction of biomaterials and cells (Fig. 2(b)) [57]. For in vitro
approaches, autologous cells, taken from patients, are seeded on
biodegradable scaffolds and matured in a bioreactor with growth
factors, mechanical stimulations, and/or signaling molecules. Such
preconditioning results in the formation of neo-tissue of heart
valves with the improved mechanical and hemodynamic properties. Easily accessible autologous cell sources, a biodegradable
scaffold with the ability to mimic the shape of heart valve, and
bioreactor processing are the essential elements for the development of in vitro TEHVs. The constructs are then implanted in animal
models and are evaluated for in vivo biological performances, such
as thrombo-resistance, in-situ tissue remodeling, and integration
with the native tissue. Following the optimization process in animal
models and assessing the biosafety aspects, the constructs can be
implanted in human clinical trials. In vitro approaches, however,
suffer from high costs, risk of infection, and lack of biomechanical
durability.
A recent development in the elds of microfabrication, tissue
engineering, and nanotechnology has revolutionized the

A. Hasan et al. / Biomaterials 103 (2016) 278e292

281

Fig. 2. Tissue engineering of heart valves: In-situ, in vitro, and hybrid approaches: a) Macroscopic picture of cell-populated (i), mesenchymal stem cells (MSCs) differentiation and
reseeding of in-situ TEHV. MSCs were differentiated into osteogenic (red) and chondrogenic (blue) cell lineages (ii), b) Photographs of decellularized whole rat heart included heart
valves (i). Staining of MHC, vWF, and connexin-43 showing small channels and gap junctions in recellularized heart (ii and iiii), c) The engineered hybrid leaet with stainless with
steel mesh core (left), and Nitinol mesh embedded within layers of cells (right). Fig (a) is reproduced from Ref. [41] and (c) from Ref. [51] with permission from Elsevier Science, and
(b) from Ref. [57] with permission from Nature Publishing Group. (For interpretation of the references to colour in this gure legend, the reader is referred to the web version of this
article.)

technology of fabricating biomimetic TEHVs and continue to play a


key role in synthesis of in vitro heart valves. For instance, 3D bioprinting technology was used to produce layer by layer accumulative products and to print cell-laden hydrogels. In addition,
micropatterning techniques with the ability to pattern specic
growth factors might be used in combination with 3D bioprinting
to promote site-specic differentiation of stem cells and progenitor
cells into VICs and VECs. FLK1 progenitor cells were shown preferentially differentiate into endothelial cells on a surface with
immobilized vascular endothelial growth factor (VEGF) and to
smooth muscle-like cells on surfaces without VEGF, arranged according to the fabricated topography and oriented by the VEGF [58].
The same technique can be implemented to engineer heart valve
tissue from stem cells taken from patients. The surface characteristics of tissue scaffolds inuence the interactions between the
biomaterial and the biological systems. The topography of the
surface can directly regulate cell functions like cell adhesion, proliferation, differentiation, and apoptosis at the interface of valve
tissue and surrounding biological environment, thus changing the
chemical and physical properties of the scaffold's surface [59]. For
instance, the broblasts can better adhere and align along specic
topographic features such as nanoscale ridges made on the gel
surface [60,61]. The alignment of endothelial cells enhances the
hemodynamic function, durability of the valve, and prevents
thrombogenicity of implanted valves [59,62].
4. Micro and nanotechnologies in HVTE
Micro and nanotechnologies play essential roles in reproducing
biomimetic anatomical 3-dimensional (3D) geometry of heart
valves to enable conduit hemodynamics, and in developing strategies to improve the function of TEHVs. Micro and nanofabrication
techniques such as lithography-based techniques, 3D bioprinting,

and microuidics have the potential to facilitate the fabrication of


viable and functional valves. Such valves, as complex biomimetic
scaffolds, undergo growth and subsequent remodeling after implantation, particularly for pediatric patients [49]. Also the micro
and nanotechnologies need to be able to maintain and appropriately stimulate viable, relevant cell types such as VICs and MSCs,
since these cells have a signicant role in remodeling valvular ECM.
An efcient tissue engineering is highly dependent on the fabrication of scaffolds because the geometry, porosity, surface characteristics, degradability, and mechanical properties affect directly
the cell adhesion, growth, and remodeling [63]. The scaffold is in
direct contact with blood and thereby it should also be resistant to
thrombosis [64]. The common micro and nanoscale technologies
employed in the synthesis of HVTEs include the micro and nanober based scaffold fabrication using electrospinning, application
of 3D cell encapsulated micro engineered porous hydrogels, use of
various biomolecules such as growth factors, gene delivery and
various micro and nanofabrication techniques for micro and nanotopography, as well as microuidic bioreactors, as discussed below.
4.1. Electrospun micro and nanobrous scaffolds in HVTEs
Electrospinning is the most commonly used technique to make
brous scaffolds as it applies to most polymers, easy to handle, and
cost-effective. It is capable of generating structural features on the
micro-to nanoscale [65], as well as the ability to introduce functionality or the modication of scaffold microstructure [66].
Different natural and synthetic polymers have been used as the
base polymer to fabricate nanobrous scaffolds for heart valve
leaets (Table 1). A natural polymer chitosan-based leaet was
fabricated by addition of chitosan bers into 3-D chitosan scaffold
(Fig. 3(a)) [67]. The tensile strength of the leaets was comparable
to those of native valve leaets in radial direction [67,68].

282

A. Hasan et al. / Biomaterials 103 (2016) 278e292

Table 1
Summary of the major natural and synthetic polymers used to develop heart valves.
Biodegradable
electrospun bers

Model system

Properties & outcome

Ref

Chitosan

Trileaet heart valve

[67]

PGA:PLLA

Pulmonary valve

PEUU

Cardiovascular
tissue sheet
transannular patch

Scaffold reinforced with chitosan bers offered biocompatibility and enhanced mechanical properties,
but were weaker compared to native valve leaets
Scaffold provided mechanical support for cells and allows deposition of ECM but loses mechanical
integrity
Achieved high cell density by concurrent electrospraying of cells and electrospinning of PEUU polymer
Patch completely degraded and was replaced with a viable endothelialized tissue and ECM. The
substituted tissue showed low brosis, low calcications and no thrombus
Tri-layered scaffold supported cell growth, ECM deposition, and anisotropic mechanical properties

[72]

PDO

[40]
[71]

PGS/PCL

Tri-layered scaffold
for heart valves

Cell-seeded hydrogels

Model system

properties & outcome

Ref

Fibrin gel

Trileaet and
bileaet valves
Vi-layered scaffold
Aortic heart
valve leaet
Heart valve leaet

Native aortic valve-like geometry and aligned patterns but poor mechanical properties

[68]

Anisotropic bending of the scaffold and control over the cell distribution throughout the scaffold
Biocompatible, easily modied physical properties, and induced elastin synthesis

[77]
[59]

High biostability and resistance to platelet adhesion for use as TEHV material

[66]

Collagen, elastin
Hyaluronan gel
POSS-PCU

Polyglycolic acid (PGA) polymer, poly(ester urethane) urea (PEUU)


and PGA/poly-L-lactide (PLLA) composites are the known synthetic
polymers that have been used to make brous TEHVs using highspeed electrospinning [69,70]. Electrospinning technique has
been extensively used to achieve the anisotropy matching to the
properties of native heart valve tissue. Pulmonary valves with the
desired mechanical properties were synthesized by aligning
smooth muscle cells within electrospun PEUU bers [71]. Incorporation of MSCs into a polydioxanone (PDO) electrospun valved
patch provided a biomimetic implantable valve with completely
degradable and non-thrombogenic properties [72]. The mechanical
properties of native valve leaets were recapitulated by an electrospun polyethylene glycol dimethacrylate (PEGdma)/Polylactic
acid (PLA) scaffold, where VICs and VECs were cultured in vitro and
behaved like a native leaet (Fig. 3(b)) [73].
The anisotropic arrangement of bers ensures proper functioning of the valves such that a very high stiffness can prevent
coaptation of the cusps, whereas a very low stiffness results in a
distortion and regurgitation [14]. An enhanced anisotropic mechanical property of TEHVs was achieved by modulating the exural stiffness of scaffold using secondary ber populations on the
microstructure of PEUU-base scaffolds. A combination of primary
and secondary bers modulated the bending moduli of the valve
construct and controlled the multi-directional anisotropy of mechanical properties [75]. Although electrospun TEHVs have relatively recapitulated the function of heart valves, dynamic scaffold
conditioning has shown capable signicantly in increasing the cell
attachment, expression of smooth muscle actin, and improving the
deposition of ECM proteins (Fig. 3(c)) [15,16].
The combination of electrospinning and other microfabrication
techniques such as rapid prototyping or layer-by-layer techniques
enables fabrication of complex geometries essential for tissue engineering of heart valves with the brous structure [17,18]. An
advanced integration of electrospinning and microfabrication
techniques was presented to make a tri-layered biomimetic model
of TEHV by assembling microfabricated poly(glycerol sebacate)
(PGS) and brous PGS/poly(-caprolactone) (PCL) electrospun
sheets. This engineered tissue supported the growth of VICs and
MSCs within the 3D structure (Fig. 3(d)) [74]. In a more advanced
model of electrospinning, microuidic-based ber spinning offered
a precise control of chemical delivery at micro-scale and enabled
creating bers with a controlled geometry and desired biological
properties. In this approach, several parallel streams of prepolymer

[74]

solutions and cell suspension were injected into chip followed by


the downstream cross-linking and formation of bers [19]. Despite
the advances of brous scaffolds for making heart valve leaets,
there exist challenges like the weak invasion of cells into the brous
scaffolds [76], and a technological constrain to make a united
construct since the valve root and cusps are made separately and
then recombined [32].

4.2. Three-dimensional cell encapsulated microporous hydrogels in


TEHVs
Hydrogels are known as effective scaffold materials for TEHVs
[68,77,78]. The proper selection of polymers in the synthesis of
TEHV hydrogel constructs is essential as it determines both the
biophysical properties and functions of TEHVs. Type I collagen and
brin are the major components of the ECM in native heart valves,
thereby they are the most commonly used biomaterials in making
TEHVs. One of the brin-based approaches for making TEHV used
the entrapment of dermal broblasts into brin gel within a micro
mold, resulted in a native aortic valve-like geometry along with
alignment patterns (Fig. 4(a)) [68]. Despite achieving an anisotropic
ber structure, the construct was much weaker compared to native
heart valves [68].
Synthetic polymers have also been used for making TEHVs,
thanks to their superior mechanical properties compared to natural
polymers [79]. A mixture of collagen and elastin was cast twice into
polytetrauoroethylene (PTFE) molds to generate a bi-layer heart
valve structure [77]. Silicone and PTFE were also used to make heart
valve leaets, but they suffered from low durability, thrombosis,
and cusp stiffening. Several groups have used biodegradable polymer composites of PLA-PGA to make the heart valve leaets and
then seeded the leaet with human broblasts and bovine aortic
endothelial cells to create a construct mimicking native valves.
Polyhydroxyalkanoate called poly-4-hydroxybutyrate (P4HB), a
bioabsorbable thermoplastic material, was added to PGA and
molded into the complex shape of valve leaet [78]. This scaffold
resulted in an enhanced ECM with the mechanical properties
resembling the native valves, although it degraded rapidly within a
month after implantation. Polyurethane (PU) is among the most
successful materials for developing cardiovascular devices
including heart valves scaffolds. It exhibits a structure with both the
hard and soft segments benecial to modulate polymer stiffness
[80]. However, it has a low biostability leading to a rapid

A. Hasan et al. / Biomaterials 103 (2016) 278e292

283

Fig. 3. Micro and nano-technologies for tissue engineering of the heart valves, a) Porous chitosan tri-leaet heart valve scaffold including chitosan hydrogel ber, b) SEM image of
VICs and VECs on electrospun scaffolds, c) The brin-based valves with exible tissue consist of three leaets continuous with a conduit wall, d) Microstructures of the fabricated
PGS scaffold with and without aligned brous layer and the cross-section of tri-layered scaffolds comprised of a PGS layer and bers. Fig (a) is reproduced from Ref. [67], (b) from
Ref. [73], (c) from Ref. [15], and (d) from Ref. [74] with permission from Elsevier Science.

degradation under oxidation or enzymatic activities [81]. The


alternative materials are Poly (ether urethane) (PEU) and
poly(carbonate-urea) urethane (PCUs) with a reasonable hydrolytic
stability for use as an implantable device [81]. Polymeric nanocomposite of PCU and polyhedral oligomericsilsesquioxanes (POSS)
nano-particles, (PCU-POSS) has also shown high biostability and
resistance to platelet adhesion for usage as a TEHV material
(Fig. 4(c)) [82,83]. Photopolymerizable hyaluronic acid (HA)
hydrogels with a modied physical properties and a low degradation rate are suitable scaffolds for heart valve tissues. Degradation
products of HA gels increase VIC proliferation and elevate matrix
production, particularly elastin [84]. In comparison with other
materials, HA is known for its stability to provide biological cues for
cell growth and differentiation [85,86]. Scaffolds made of synthetic
polymers within micro molds of valve conduit may generate toxic
solvents that should be removed prior to cells seeding [87]. In
addition, the end products from the degrading acidic polymers may
alter pH and causes an undesired cell death and differentiation [88].
Methacrylated HA in combination with collagen was utilized to
synthesize TEHV, which provided a brous structure of native valve

leaets. Collagen supplied increased mechanical properties,


whereas HA promoted cell proliferation [89].
A number of research groups have attempted to use various
microfabrication techniques to make either 3D molds for forming
hydrogels or bioprinting cell-laden hydrogels (Fig. 4(b,d)) [90e92].
Soft lithography is a microfabrication technique broadly used to
create 3D constructs [93]. However, attempts to engineer heart
valve conduits (integrated root and leaets) have been limited by
homogeneous material biomechanics due to the usage of single
material formulation for both the valve roots and leaets. An
improvement soft lithography technique is the direct-write printing of pre-cured hydrogels extruded using the pressure-driven
method on an XYZ stage. This permits a controlled deposition of
3D layers of polymers containing cells or bioactive materials [94].
To engineer bioprinted tissues, cells can be either embedded within
biologically-relevant hydrogels or printed free of scaffold support
[95,96]. 3D printing was used to fabricate TEHVs by using a lasersintering 3D printer to pattern a printable polymer and seed
vascular cells within the polymer. Hybrid hydrogels of methacrylated hyaluronic acid (HeMA) and methacrylated gelatin (GelMA)

284

A. Hasan et al. / Biomaterials 103 (2016) 278e292

Fig. 4. Application of micro and nano-composites and gene delivery techniques for the synthesis of TEHVs. Examples of TEHVs: a) Fibrin-based bi-leaet valve construct after one
month incubation, with cusp-like leaets attached to a cylindrical root, b) Bioprinting a heart valve conduit with the encapsulation of HAVIC within the leaets after 7-day incubation in culture tube, c) The prototype of POSSePCU nanocomposite heart valve, d) Adenovirus-encapsulated bers for the controlled gene delivery in tissue engineered
scaffolds, e) Heart valve leaet gene delivery using polyurethane (PU) pulmonary replacement cusps with antibody-tethered AdGFP. Fig (a) is reproduced from Ref. [68] with
permission from Mary Ann Liebert. Fig (b) is reproduced from Ref. [90], (c) from Ref. [83], (d) from Ref. [91] with permission from Elsevier Science. Fig (e) is reproduced from
Ref. [92] with permission from Nature Publishing Group.

were used to fabricate tri-leaet heart valve leaet with a high cell
viability and matrix remodeling ability [97]. 3D printing was also
used to print aortic valve PEG-DA scaffolds seeded with interstitial
cells and cultured for up to 21 days with 100% cells viability [98].
Synthetic living alginate/gelatin hydrogel-based valve conduits
were developed with anatomical architecture and incorporation of
dual cell types of aortic root sinus smooth muscle cells (SMCs) and
aortic leaet VICs. Aortic valve conduits were successfully bioprinted with direct encapsulation of SMC in the valve root and VIC
in the leaets, resulted in 3D heart valves made by self-assembly of
cell types constituting the valve [97]. Despite a slight drop in the
ultimate strength and peak strain of 3D bioprinted hydrogel in 7
days culture, the tensile biomechanical properties were preserved,
and a robust living tri-leaet heart valve with multiple cell populations, high cell viability, and reasonable phenotypes in culture
was developed. The cells used in bioprinted cell-laden hydrogels
need a better migration into the inner structure of the polymeric
layer [99]. In spite of potential application of bioprinting approaches for the synthesis of TEHVs, the applicability of printable
and biocompatible hydrogel materials and certainty on human
valve's cell response under heavy biomechanical loading are still
questionable. The achievement of anisotropic mechanical properties of native valve is a challenge for 3D bioprinting techniques.
However, the combination of 3D bioprinters and electrospinning
has a unique advantage to improve the anisotropy of 3D hydrogels.
The fabrication of alternative layers of electrospun PCL and
chondrocyte-laden brin/collagen hydrogel layer is an example of
using this combinatory technique [100].
4.3. Gene delivery, growth factors, and biomolecules in HVTE
Micro and nano-technologies can play signicant roles to the
synthesis of TEHVs in many ways including controlled delivery of
growth factors as well as gene therapy for regaining functional
heart valve tissues.

4.3.1. Gene delivery


Gene delivery methods have been developed to introduce specic foreign DNA pieces into a hosting cell for the purposes like
expression of special proteins [101]. In general, three main gene
delivery techniques have been used as: physical means like electroporation [102], ultrasound, and hydrodynamics, via viral carriers
(usually adenoviruses and retroviruses) [103e105], or non-viral
gene carriers. Although physical and viral vector methods present
high efciency (as high as 80e90% for viral delivery and 50e70% for
electroporation), there is a risk for insertion of viral carrier's nucleic
acid into the hosting cell with unfavorable subsequences or cell
damages by applying physical stimulations [101,105]. On the other
hand, although the transfection methods using non-viral carriers
have shown a low efciency (20e30%), the method is safe enough
for the clinical purposes [102]. Thus, non-viral methods show a
high potential in gene delivery for biomedical applications and are
under investigation to enhance the efciency of transfection [101].
Novel methods have been recently developed taking advantage
of nano/micro technologies using non-viral gene carriers. These
carriers are usually based on lipids, polymers, and nanostructures
[106e110]. Application of nanostructures has been drastically
increasing. Because of the small size of nanoparticles, they can
enter the cells and thereby are considered as valuable biomedical
tools for different applications like therapeutics and gene delivery.
They have shown promising potentials in the delivery of genes into
cells by nanoparticles functionalized with DNA pieces [111e123].
Magnetic nanoparticles (MNPs) [111e115,124,125], dendrimers
[116e121,126e129], and exosomes [122,123,130] are the most
popular nanoparticles that have attracted attentions during the last
decade. MNPs are more advantageous compared to other nonvirals due to the higher cell viability, less needs to transfection
materials, higher efciency, shorter transfection time, and minimal
effects on cell differentiation and proliferation [111e115,124,125].
Reports indicate that among magnetic materials, the presence of
compounds like CoFe2O4, NiFe2O4, and MnFe2O4, with a high

A. Hasan et al. / Biomaterials 103 (2016) 278e292

magnetic moment in the nanoparticle increases the efciency of


the transfection [131,132]. Dealing with these nanoparticles, one
should always take into account their high toxicity to cells
[133e136]. However, among the magnetic materials, compounds
like maghemite (g-Fe2O3) and magnetite (Fe3O4) are relatively safe,
whereas they have high magnetic moments [137e139]. Therefore,
magnetite is a widely used compound for MNPs for biomedical
applications [131,132,137e139]. Nevertheless, for biomedical applications, even in the case of using biologically safe compounds,
the surface of MNPs should be modied [140] and coated with
compounds like synthetic organic polymers (polyethylene glycol),
polyvinyl alcohol, poly-L-lactic acid) [141e143], gold [140,144], or
natural polymers (proteins and carbohydrates) [141,145e148], to
improve their functionality and prevent their agglomeration and
cytotoxicity [149,150]. Coating the MNPs with some compounds
like polyethylenimine (PEI), not only decreases dramatically the
cytotoxicity [150], but also acts as a dispersant [151] and enhances
the transfection efciency [152e155]. Some nanoparticles are,
however, shown to cause calcication of heart valves [156]. New
advances in nanocomposite materials such as (polyhedral oligomeric silsesquioxaneepoly(carbonate-urea)urethane; POSSePCU)
nanocomposites [157] or PCU-based valves coated with POSS
nanoparticles [158] have shown promising results to avoid calcication and are expected to be tested for the future cardiovascular
tissues and TEHVs [82,83].
For gene delivery into the heart, naked plasmids [159],
DNAelipid complexes [160], DNAepolymer complexes [161],
lentivirus [162e164], and adenovirus [92] (Fig. 4(e)) are the most
applied non-viral and viral methods. Gene delivery to cardiovascular systems using non-viral methods is usually limited by a low
efciency of transfection [165,166]. Using nanostructures, as nonviral carriers, has shown promising results to overcome this limitation [165e167]. Recently, more attention has been attracted to
the gene delivery to heart tissues using nanostructures like MNs
[165,167]. However, more studies are needed to address advantages
and disadvantages of gene delivery techniques based on micro and
nanostructures for the articial heart and to explore the whole span
that are safe and functional candidates for gene delivery.
4.3.2. Growth factors
The requirement of cell seeding prior implantation is considered
as an important limitation for TEHVs' application. Inducing recellularization using growth factors is an alternative strategy that
helps to overcome the cell seeding limitation [168]. One strategy is
to coat the scaffold with growth factors [169]. A number of growth
factors such as vascular endothelial growth factor (VEGF), hepatocyte growth factor (HGF), broblast growth factor (FGF), stromal
cell-derived growth factor-1 (SDF-1), connective tissue growth
factor (CTGF), hepatoma-derived growth factor (HDGF), transforming growth factor-b1 (TGF-b1), basic broblast growth factor
(bFGF), and platelet-derived growth factor (PDGF) [169e178] have
been shown to be useful for tissue engineering of cardiovascular
systems. In VSMCs, TGFb led to a dramatic increase in matrix production [179]. PDGF accelerated VSMCs proliferation in a matrix of
collagen [180]. Incorporation of basic broblast growth factor
(bFGF), resulted in an increment in collagen synthesis [170]. It was
also shown that a close collaboration between ECM and growth
factors is essential for a successful tissue engineering of heart valves
with a stable cell-matrix adhesion, reconstitution, and reendothelialization [169,181]. For instance, a combination of hepatocyte growth factor and bronectin had a synergetic effect in the
formation of tissue engineered heart valves [181]. Furthermore,
growth factors can lead to multiple favorable impacts on TEHV. For
example, TGFb is an important factor not only for cell proliferation
but also for ECM growth in TEHV [182]. Heparin-VEGF has been

285

used to coat the decellularized aortic heart valves with the purpose
of acting as antithrombotic agent, whereas a-SMA expression using
growth factor-h1 (TGF-h1) on valvular interstitial cells induced the
migration of endothelial progenitor cells on decellularized valves
[89].
The modication of nanoscale morphology of the scaffold surface increases the surface area to volume and provides nanoscale
drug carrier coatings for the use in TEHVs [90]. Also a majority of
TEHVs perform satisfactorily within an in vitro microenvironment,
but they do not depict a strong function upon implantation.
Microfabricated organ-on-chip technology has recently been
introduced to provide platforms for modeling the in vivo microenvironment of different organs and particularly to assess the inuence of drugs, genes, and biomolecules on the function of cells
and tissues. Microuidics was used to provide a pulsatile ow of
physiologic systole and diastolic phase within bioreactors and to
supply transvalvular pressure and ow conditions [183]. Cells
seeded within a microuidic bioreactor for the synthesis of a trileaet construct showed an improved cell proliferation, ECM synthesis, and led to enhanced tensile properties with respect to the
native pulmonary valves [184]. Also a microuidic-based heart
valve-on-chip was developed to investigate the impact of shear
stress-regulated paracrine interactions between VECs and VICs as
well as studying the diverse aspects of gene delivery mechanism
and vascular/valvular biology for modeling the in vivo performance
of heart valves [185].
5. Application of computational modeling in heart valve
tissue engineering
The recent advances in high-performance computing and
noninvasive imaging have made possible the numerical simulation
of complex biological systems, such as heart valves. As a result,
researchers have gained a better understanding of the heart valves'
biomechanics via designing effective prosthetic valves, and developing novel diagnostic tools, treatment modalities, and surgical
procedures for heart valve diseases [186e188].
Despite the ubiquity of computation in many disciplines, its
application in tissue engineering is still relatively nascent
[189e192], particularly for heart valves. This is partly due to a fast
progress in technologies that have outpaced computational
research. Numerical simulations (referred herein as simulation) of
biological systems, whether at the organ, tissue, cellular or molecular scales, can be extremely computationally intensive. Therefore, development of techniques that take advantage of modern
hardware are required both at the algorithmic and scientic levels
(e.g., novel mathematical formulations).
Simulation of tissue-engineered heart valves as a complement
to the experimental models can be categorized as 1) macroscale
computation and 2) molecular/nanomicroscale computation. The
former is concerned with the mechanical, thermal, and electrical
properties of valve tissues, whereas the latter is concerned with
properties and composition of valve tissues at the micro and nanostructure level [193]. Continuum modeling (nite element analysis
(FEA) [194], computational uid dynamics (CFD) [195]), and molecular dynamics (MD)-based models are the models typically used
in these simulations [196].
5.1. Macroscale modeling: tissue mechanical properties and design/
evaluation of bioprosthetic valves
Macroscale models of heart valves in general use uid-structure
interaction (FSI), which have been developed extensively over the
years for heart valve systems [197] and applied to bioprosthetic
valves [198e204]. Such models allow researchers to simulate

286

A. Hasan et al. / Biomaterials 103 (2016) 278e292

complex in vivo conditions, perform parametric studies, and


consider novel materials [158,205]. Computer simulation provides
a cost-effective alternative to experimentation that can be costly,
impractical or impossible, may be difcult to repeat, and in general,
does not capture the true in vivo conditions.
Many studies have been conducted to characterize mechanical
heart valves [206], which have been known to induce thromboembolic events due to large non-physiologic ows [207,208]. Bioprosthetic valves have shown signicantly less thrombogenecity,
but they suffer from durability issues such as stiffening due to
calcication [209]. Thromboegenicity of heart valves has been
linked to platelet activation in the blood due to large shear stresses
from the non-physiologic ows. Using suspended-uid solid
computational uid dynamics (CFD) models, it is possible to
simulate thromboembolic events and engineer more effective
bioprosthetic heart valves [210e212].
One of the key issues in the simulation of heart valves is to
capture material responses of valve leaet tissues which are
inherently nonlinear and anisotropic. As explained in section 4,
these biomechanical properties are a consequence of the collagen
structure and alignment at the cellular and molecular scales [213], a
result of contractile forces generated by cells [214e217]. It is
possible to accurately capture tissue stress generation and heterogeneous collagen arrangement in engineered tissues [218e223].
Modern techniques allow for the direct determination of material properties by explicitly modeling the microstructure and
simultaneous modeling the effect on cells (Fig. 5(a)) [224e226].
Such techniques are referred to as volume averaging or homogenization methods and involve explicitly modeling a so-called
representative volume element (RVE) of the microstructure. It has
been observed that a large portion of the anterior leaet contains
cardiac muscle bers. Valve leaets are typically considered as
passive elements, but passive material models have tended to
result in non-physiological bulging during systole [227]. By accounting for the activation of muscle bers in mitral valve leaets,
bulging effect can be reduced, indicating that the valve leaets are
not fully passive [228]. This indicates that the muscle bers must be
accounted for biomimetic engineering of valve replacements.
5.2. Micro/nanoscale modeling
Micro and nanoscale computational models are applied to
characterize the structural property relationship at the micro and
nanoscale, and to design the tissue-engineering scaffolds. Finite
element (FE) models are useful to optimize the shape, structures,
and compositions of scaffolds (Fig. 5(c,d)) [230]. Simple numerical
schemes can also be used to study, for instance, the diffusion of
oxygen and nutrients through the engineered valve [231]. The
lattice-Boltzmann method (LBM) has been employed to quantify
ow-induced shear stress in porous scaffolds, overcoming the
severely limited analytical techniques classically used [189,232]. A
slightly different approach involves digital volume correlation
(DVC) with an inverse FEM modeling to compute properties of
scaffolds [233].
Multiscale methods have been employed to design and predict
properties of scaffolds and account for the microstructural features
of techniques such as electrospinning (Fig. 5(b)) [229,234]. The
electronic structure of nanoscale particles, such as hydroxyapatite
(HAp), is obtained using quantum mechanical computations. The
stiffness can then be computed by applying strains to unit cells and
computing the total energy. This stress-strain data is then used for
the RVE of the scaffold in the homogenization procedure. Applications to heart valve tissue are still, however, work in progress
[225,226].
There has been relatively little works in the nanomechanical

modeling of the brillar collagen network and the negatively


charged GAG side chains of aggrecan, responsible for the biomechanical properties of tissues. Different nanomechanical and
poroelastic models have been developed to study the effect of
poroelasticity and viscoelasticity of the ECM mechanisms on the
energy dissipation and time-dependent behavior of cartilage, and
to investigate the ber-dependent behavior of soil mechanics
[235,236]. However, the application of these techniques to tissueengineered valve tissue is essential for modeling the desired mechanical properties. This is particularly important for the long-term
properties such as degradation.
Interstitial growth models have been employed heavily to study,
for instance, atherosclerosis [237], but the application to tissueengineered valves has yet to be done. Ateshian et al. [238] developed a mixture-theory based framework that can be used to model
growth and to further combined with FE models. Another framework is to use RVE volume averaging method to compute nutrient
diffusion at the microscale and obtain the macroscopic growth of
tissue scaffolds [239].
In addition to the large-scale modeling of heart valves, cellular
mechanotransduction (the science of characterizing the cells
respond to mechanical stimuli in different cellular processes) has a
high potential for small scale modeling of TEHVs. These processes
underlie morphogenesis, differentiation, determination, development, and pathology. It also can investigate the role of mechanical
cues on stem cell differentiations [240] and generally be used to
nd out the tissue stiffness that most resembles that of the substrate [241]. Being able to simulate this can be the key to engineering biomimetic valve tissues [242].
6. Challenges and future work
In overall, one primary challenge of in-situ heart valve models is
to recruit appropriate cells into the scaffold to further develop heart
valve tissue into the body. The existing in-situ scaffolds are not
smart enough to provide selective mechanisms for recruitment of
target circulating cells and to regulate their behavior for deposition
of anisotropic load-bearing scaffold and formation of a loadbearing tissue. This drawback necessitates controlled patterning
and functionalization of scaffolds with biological materials,
biochemical modulators, and growth factors for specic recruitment of cells. Micro and nano-technologies have been employed
into in vitro models of heart valves, however their in-situ performance in supporting the production of biochemical gradients and
anisotrpic biomechanical characteristics resulting in selecting the
migration of progenitor cells has not been reported yet. Also given
the complex hemodynamic loads on HVs and the rapid altration in
their deformation patterns, the recruited cells are exposed to
altering biophysical cues that impact the cell fate. There is a crucial
need for the development of novel biomimetic high-throughput
model systems to investigate systemically the effects of loading
cycles, the biomechanical anisotropy of scaffolds, gender variations,
and cell types on recruitment of cells into in-situ scaffolds. Development of such powerful tools is instrumental for achieving
implantable in-situ heart valves, to model the diseases associated
with heart valves, and to predict their human in vivo performance.
Bioprinting as an emerging technique for printing biological
tissues and cell-laden scaffolds is unique to produce tissues with
diverse morphological, mechanical, and biochemical features.
These capabilities have enabled fabrication of 3D heart valve geometries in vitro with the mechanical heterogeneity of native valves
with a potential application for translational implantation. Despite
successful bioprinting of hydrogel-based heart valves discussed in
section 3, the stability and biomimetic mechanical characteristics of
these constructs under mechanical loading for a prolonged time are

A. Hasan et al. / Biomaterials 103 (2016) 278e292

287

Fig. 5. Computational modeling of TEHVs, a) Representative volume element (RVE) of valve tissue [225], b) RVE of scaffold model [229], c) Leaet stresses with ber orientation of
engineered tissue with orthotropic scaffold [230], d) Nanoscale poroelastic indentation. Fig (a) is reproduced from Ref. [225] with permission from American Society of Mechanical
Engineering. Fig (b) is reproduced from Ref. [229], (c) from Ref. [230], (d) from Ref. [230] with permission from Elsevier Science.

questionable. Current bioprinting systems have not been able to


print implantable heart valves after in vitro culturing. The existing
3D bioprinted valves are too soft to withstand hemodynamic mechanical loading, thereby they cannot recapitulate the mechanical
durability of native heart valves. Further works for translation of
the bioprinting strategy to the clinic need a better understanding of
the scaffold remodeling under simulated hemodynamic conditions,
and incorporation of multiple polymer blends into heterogeneous
valve conduits to mimic both low and high strain behaviors of valve
tissues. New biomaterials cross-linked chemically or through UV
light exposure need to be developed to balance the ability to
withstand biomimetic mechanical pressure and to provide an
appropriate microenvironment for cells growth, function, and
survival [243,244].
While TEHVs exposed to hemodynamic mechanical loading can
provide insight into heart valve function in the body, there is still
needs to develop non-destructive and real-time systems for postimplantation monitoring of cells inltration, and scaffolds degradation and remodeling. The conventional histological analysis requires terminating the heart valve function and cannot record the
time-dependent feature of the remodeling. Echo/doppler techniques and uorescence imaging techniques are the candidates for
monitoring the dynamic recruitment and activity of cells and to
record the scaffold growth.
Detailed investigation is still required into the basic science of
inammatory response in TEHVs. In this regards, the in vitro TEHVs
models produced from synthetic brous scaffolds are preferable

respect to the decellularized tissues given the fact that they are
synthesized from materials that are essential for cell adhesion and
function with minimally induced immune response. However,
technological restrains in fabrication of 3D biomimetic scaffolds
limits their translation to the clinical use. There are evident challenges associated with the storage and establishment of quality
control standards. Incorporation of bioactive materials for the
controlled recruitment of cells and regulating the biochemical
properties of ECM can also lead to new challenges for biosafety,
inammation, thrombosis, angiogenesis, and efcacy performance.
A key strategy for the TEHVs studies should be identifying specic
signaling pathways that are able to regulate post-infarction
inammation, limit scaffold degradation, and manage tissue
remodeling and scarring.
When considering tissue-engineering applications, it is of
utmost importance to consider the micro and nanoscale mechanics
and features. Even with the rapidly increasing technology, experimentation at the micro-and nanoscales is costly, difcult, and often
impractical. Computational techniques provide a very powerful
method to circumvent these limitations. Although the high cost and
unavailability of fresh human heart valves is an obstacle in
obtaining extensive biomechanical data of heart valves, a combined
experimental-computational modeling approach can provide
extensive insights into structure-biomechanical response relationships from a small number of heart valve samples. Specifically, extensive parametric studies using computational models
that have been validated against simple experimental test data will

288

A. Hasan et al. / Biomaterials 103 (2016) 278e292

be helpful in design and development of functional TEHV


constructs.
The state-of-the-art in computational modeling of tissueengineering is to use noninvasive imaging techniques, such as
MRI and CT, to reconstruct accurate 3D models of scaffolds and
tissue and to use multiscale and multiphysics numerical simulation
techniques, such as FEM and CFD to understand their behavior.
Such models can be used to predict and design articial tissues that
mimic the natural heart valve structure and behavior for valve replacements. The key lies in multiscale and multiphysics modeling,
particularly, processes at the nanoscale. Future studies should
consider 1) particle transport (nutrients, oxygen, etc), 2) growth
models, 3) coupling nanoscale processes with macroscopic models,
4) mechanotransduction, 5) patient-specic intra-operative models
for surgical implantation of bioprosthetic valve replacements, and
6) design of prosthetic valves that mimic natural behavior and use
novel materials.
Last but not least, dening the most relevant biomechanical
properties crucial to the success of functional and implantable
TEHV constructs, as well as straightforward experiments in which
these properties can be determined are important factors that
deserve more attention in future works. Similarly, the appropriate
ranges of the scaffold and tissue construct properties required for
acceptable functionality are also among important factors
deserving attention. Ensuring sufcient biomechanical properties
in the TEHVs will be important prior to clinical trials. This requires
designing new biomaterials with improved nano-microscale features as well as improved microenvironment and bioreactors for
better preconditioning of the TEHV constructs.

7. Conclusion
Tissue-engineered heart valves are currently the sole technology to create physiologically relevant tissues analogous to native
human valves with fewer side effects and longer sustainability for
future valve replacement. The challenges of strength and durability
under hemodynamic mechanical loading and rejection of explanted valves due to the inammatory response of host tissue are the
bottleneck of these engineering heart valves and yet to be tackled.
In this review, we addressed the advantages of micro and nanotechnology for self-organizing the cells behavior according to the
guidance cues provided by the ECM and for recapitulating the
structure and function of in vitro and in-situ tissue engineered heart
valves. Micro and nano-technologies have signicantly helped to
engineer topography of scaffolds, delivery of molecules and cells to
the specic site of tissues, and complex geometries and multidirectional anisotropic properties of valve scaffolds. This technology can be further linked to novel bioreactors to advance biomechanical properties of engineered valves as a key step for
translating the TEHVs to implantable valves, speeding up the process of treatment and allowing patients to receive cardiac
treatment.
Selection of appropriate scaffolds, choice of cell seeding method,
understanding remodeling of the scaffold seeded with cells, design
of biomimetic bioreactors for process conditioning and testing
synthesized valves for further scale-up operation, incorporation of
effective anti-inammatory drugs modulating the immune
response, and adjustment of waiting time for seeding the cells onto
scaffolds to provide of the shelf TEHVs, accommodating patientto-patient heterogeneity via considering the design aspects, and
predicting function of explanted valves as early as possible are
critical future steps for successful development of long-lasting, selfrepairing, functional, and implantable heart valves.

Acknowledgement
The authors acknowledge the Natural Sciences and Engineering
Research of Canada, NPRP9-144-3-021 from Qatar Foundation,
QUUG-CENG-MIE-15/16-7 and QUST-CENG-FALL-15/16-20 from
Qatar University, the Farouk Jabre interdisciplinary research award
from American University of Beirut, and the CNRS grant from National Council for Scientic Research, Lebanon, for the support for
this paper.
References
[1] D.T. Simionescu, J. Chen, M. Jaeggli, B. Wang, J. Liao, Form follows function:
advances in trilayered structure replication for aortic heart valve tissue engineering, J. Healthc. Eng. 3 (2012) 179e202.
[2] M.K.S. Loftin, Y.W. Chun, A. Khademhosseini, W.D. Merryman, EMT-inducing
biomaterials for heart valve engineering: taking cues from developmental
biology, J. Cardiovasc. Trans. Res. 4 (2011) 658e671.
[3] M.H. Yacoub, J.J.M. Takkenberg, Will heart valve tissue engineering change
the world? Nat. Clin. Pract. Cardiovasc. Med. 2 (2005) 60e61.
[4] J.M. Reimer, Z.H. Syedain, B.H. Haynie, R.T. Tranquillo, Pediatric tubular
pulmonary heart valve from decellularized engineered tissue tubes, Biomaterials 62 (2015) 88e94.
[5] C.A. Durst, M.P. Cuchiara, E.G. Manseld, J.L. West, K.J. Grande-Allen, Flexural
characterization of cell encapsulated PEGDA hydrogels with applications for
tissue engineered heart valves, Acta Biomater. 7 (2011) 2467e2476.
[6] I. Vesely, Heart valve tissue engineering, Circulat. Res. 97 (2005) 743e755.
[7] A. Khademhosseini, J.P. Vacanti, R. Langer, Progress in tissue engineering, Sci.
Am. Mag. 2009 (May 2009).
[8] Y. Du, E. Lo, S. Ali, A. Khademhosseini, Directed assembly of cell-laden
microgels for fabrication of 3D tissue constructs, PNAS 105 (2008)
9522e9527.
[9] T.C. Flanagan, A. Pandit, Living articial heart valve alternatives: a review,
Eur. Cells Mater. 6 (2003) 28e45.
[10] R. Gauvin, M. Guillemette, R. Langer, A. Khademhosseini, Emerging trends in
tissue engineering, in: Zhanfeng Cui (Ed.), Comprehensive Biotechnology,
Elsevier Ltd, 2011.
n, A. Driessen-Mol, C. Bouten, F. Baaijens, J.L. de la
[11] D. MacGrogan, G. Luxa
Pompa, How to make a heart valve: from embryonic development to
bioengineering of living valve substitutes, Cold Spring Harb. Perspect. Med. 4
(2014) a013912.
[12] C.V. Bouten, A. Driessen-Mol, F.P. Baaijens, In situ heart valve tissue engineering: simple devices, smart materials, complex knowledge, Expert Rev.
Med. Devices 9 (2012) 453e455.
[13] N.J.B. Driessen, A. Mol, C.V.C. Bouten, F.P.T. Baaijens, Modeling the mechanics
of tissue-engineered human heart valve leaets, J. Biomech. 40 (2007)
325e334.
[14] J.-H. Chen, C.A. Simmons, Cellematrix interactions in the pathobiology of
calcic aortic valve disease critical roles for matricellular, matricrine, and
matrix mechanics cues, Circulat. Res. 108 (2011) 1510e1524.
[15] T.C. Flanagan, C. Cornelissen, S. Koch, B. Tschoeke, J.S. Sachweh, T. SchmitzRode, et al., The in vitro development of autologous brin-based tissueengineered heart valves through optimised dynamic conditioning, Biomaterials 28 (2007) 3388e3397.
[16] H. Hong, N. Dong, J. Shi, S. Chen, C. Guo, P. Hu, et al., Fabrication of a novel
hybrid scaffold for tissue engineered heart valve, J. Huazhong Univ. Sci.
Technol. Med. Sci. 29 (2009) 599e603.
[17] P.M. Dohmen, W. Konertz, Tissue-engineered heart valve scaffolds, Ann.
Thorac. Cardiovasc. Surg. Off. J. Assoc. Thorac. Cardiovasc. Surg. Asia 15
(2009) 362e367.
[18] M. Van Lieshout, C. Vaz, M. Rutten, G. Peters, F. Baaijens, Electrospinning
versus knitting: two scaffolds for tissue engineering of the aortic valve,
J. Biomaterials Sci. Polym. Ed. 17 (2006) 77e89.
[19] H. Onoe, T. Okitsu, A. Itou, M. Kato-Negishi, R. Gojo, D. Kiriya, et al., Metrelong cell-laden microbres exhibit tissue morphologies and functions, Nat.
Mater. 12 (2013) 584e590.
[20] M. Misfeld, H.H. Sievers, Heart valve macro- and microstructure, Philos.
Trans. R. Soc. B Biol. Sci. 362 (2007) 1421e1436.
[21] M.A.J. Cox, Local Mechanical Properties of Tissue Engineered Heart Valves,
Technische Universiteit Eindhoven/Eindhoven University of Technology,
2009.
[22] E.O. Carew, J. Patel, A. Garg, P. Houghtaling, E. Blackstone, I. Vesely, Effect of
specimen size and aspect ratio on the tensile properties of porcine aortic
valve tissues, Ann. Biomed. Eng. 31 (2003) 526e535.
[23] J.T. Butcher, G.J. Mahler, L.A. Hockaday, Aortic valve disease and treatment:
the need for naturally engineered solutions, Adv. Drug Deliv. Rev. 63 (2011)
242e268.
[24] T.C. Flanagan, A. Pandit, Living articial heart valve alternatives: a review,
Eur. Cells Mater. 6 (2003) 28e45.
[25] M.S. Sacks, W.D. Merryman, D.E. Schmidt, On the biomechanics of heart
valve function, J. Biomech. 42 (2009) 1804e1824.

A. Hasan et al. / Biomaterials 103 (2016) 278e292


[26] R.B. Hinton, K.E. Yutzey, Heart valve structure and function in development
and disease, Annu. Rev. Physiol. 73 (2011) 29.
[27] N.J. Brand, A. Roy, G. Hoare, A. Chester, M.H. Yacoub, Cultured interstitial
cells from human heart valves express both specic skeletal muscle and nonmuscle markers, Int. J. Biochem. Cell Biol. 38 (2006) 30e42.
[28] A. Hasan, A. Paul, N.E. Vrana, X. Zhao, A. Memic, Y.-S. Hwang, et al., Microuidic techniques for development of 3D vascularized tissue, Biomaterials 35
(2014) 7308e7325.
thy, T. P
[29] S. Deme, I. Apa
azm
andi, E. Benton, G. Reitz, Y. Akatov, On-board TLD
measurements on MIR and ISS, Radiat. Prot. Dosim. 120 (2006) 438e441.
[30] A. Hasan, A. Memic, N. Annabi, M. Hossain, A. Paul, M.R. Dokmeci, et al.,
Electrospun scaffolds for tissue engineering of vascular grafts, Acta Biomater.
10 (2014) 11e25.
[31] D.Y. Cheung, B. Duan, J.T. Butcher, Current progress in tissue engineering of
heart valves: multiscale problems, multiscale solutions, Expert Opin. Biol.
Ther. 15 (2015) 1155e1172.
[32] F.W. Sutherland, T.E. Perry, Y. Yu, M.C. Sherwood, E. Rabkin, Y. Masuda, et al.,
From stem cells to viable autologous semilunar heart valve, Circulation 111
(2005) 2783e2791.
[33] J.T. Butcher, G.J. Mahler, L.A. Hockaday, Aortic valve disease and treatment:
the need for naturally engineered solutions, Adv. Drug Deliv. Rev. 63 (2011)
242e268.
[34] M.S. Sacks, F.J. Schoen, J.E. Mayer Jr., Bioengineering challenges for heart
valve tissue engineering, Annu. Rev. Biomed. Eng. 11 (2009) 289e313.
[35] C.H. Chang, C.C. Chen, C.H. Liao, F.H. Lin, Y.M. Hsu, H.W. Fang, Human acellular cartilage matrix powders as a biological scaffold for cartilage tissue
engineering with synovium-derived mesenchymal stem cells, J. Biomed.
Mater. Res. Part A 102 (2014) 2248e2257.
[36] A. Sanz-Garcia, J. Oliver-De-La-Cruz, V. Mirabet, C. Ganda, A. Villagrasa,
E. Sodupe, et al., Heart valve tissue engineering: how far is the bedside from
the bench? Expert Rev. Mol. Med. 17 (2015) e16.
[37] D. Schmidt, P.E. Dijkman, A. Driessen-Mol, R. Stenger, C. Mariani, A. Puolakka,
et al., Minimally-invasive implantation of living tissue engineered heart
valves: a comprehensive approach from autologous vascular cells to stem
cells, J. Am. Coll. Cardiol. 56 (2010) 510e520.
[38] D. Van Geemen, A. Driessen-Mol, L.G. Grootzwagers, R.S. Soekhradj-Soechit,
P.W. Riem Vis, F.P. Baaijens, et al., Variation in tissue outcome of ovine and
human engineered heart valve constructs: relevance for tissue engineering,
Regen. Med. 7 (2012) 59e70.
[39] A. Mol, A.I. Smits, C.V. Bouten, F.P. Baaijens, Tissue engineering of heart
valves: advances and current challenges, Expert Rev. Med. Devices 6 (2009)
259e275.
[40] J. Puvimanasinghe, J. Takkenberg, M. Edwards, M. Eijkemans, E. Steyerberg,
L. Van Herwerden, et al., Comparison of outcomes after aortic valve
replacement with a mechanical valve or a bioprosthesis using microsimulation, Heart 90 (2004) 1172e1178.
[41] P.E. Dijkman, A. Driessen-Mol, L. Frese, S.P. Hoerstrup, F.P. Baaijens, Decellularized homologous tissue-engineered heart valves as off-the-shelf alternatives to xeno-and homografts, Biomaterials 33 (2012) 4545e4554.
[42] C.D. Daly, G.R. Campbell, P.J. Walker, J.H. Campbell, In vivo engineering of
blood vessels, Front. Biosci. 9 (2004) 1915e1924.
[43] M. Yamanami, Y. Yahata, T. Tajikawa, K. Ohba, T. Watanabe, K. Kanda, et al.,
Preparation of in-vivo tissue-engineered valved conduit with the sinus of
Valsalva (type IV biovalve), J. Artif. Organs 13 (2010) 106e112.
[44] M. Yamanami, Y. Yahata, M. Uechi, M. Fujiwara, H. Ishibashi-Ueda, K. Kanda,
et al., Development of a completely autologous valved conduit with the sinus
of valsalva using in-body tissue architecture technology a pilot study in
pulmonary valve replacement in a beagle model, Circulation 122 (2010)
S100eS106.
[45] D. Kehl, B. Weber, S. Hoerstrup, Bioengineered living cardiac and venous
valve replacements: current status and future prospects, Cardiovasc. Pathol.
25 (2016) 300e305.
[46] S.P. Nejad, M.C. Blaser, J.P. Santerre, C.A. Caldarone, C.A. Simmons, Biomechanical conditioning of tissue engineered heart valves: too much of a good
thing? Adv. Drug Deliv. Rev. 96 (2016) 161e175.
[47] J. Bechtel, M. Mller-Steinhardt, C. Schmidtke, A. Brunswik, U. Stierle,
H. Sievers, Evaluation of the decellularized pulmonary valve homograft
(SynerGraft), J. Heart Valve Dis. 12 (2003), 734-9; discussion 9e40.
[48] P. Mathieu, R. Bouchareb, M.-C. Boulanger, Innate and adaptive immunity in
calcic aortic valve disease, J. Immunol. Res. 2015 (2015).
[49] M. Sewell-Loftin, Y.W. Chun, A. Khademhosseini, W.D. Merryman, EMTinducing biomaterials for heart valve engineering: taking cues from developmental biology, J. Cardiovasc. Transl. Res. 4 (2011) 658e671.
[50] N. Grabow, K. Schmohl, A. Khosravi, M. Philipp, M. Scharfschwerdt, B. Graf, et
al., Mechanical and structural properties of a novel hybrid heart valve scaffold for tissue engineering, Artif. Organs 28 (2004) 971e979.
[51] S.H. Alavi, W.F. Liu, A. Kheradvar, Inammatory response assessment of a
hybrid tissue-engineered heart valve leaet, Ann. Biomed. Eng. 41 (2013)
316e326.
[52] G. Steinhoff, U. Stock, N. Karim, H. Mertsching, A. Timke, R.R. Meliss, et al.,
Tissue engineering of pulmonary heart valves on allogenic acellular matrix
conduits in vivo restoration of valve tissue, Circulation 102 (2000). Iii-50-Iii5.
[53] H. Hong, N. Dong, J. Shi, S. Chen, C. Guo, P. Hu, et al., Fabrication of a novel
hybrid heart valve leaet for tissue engineering: an in vitro study, Artif.

289

Organs 33 (2009) 554e558.


[54] M. Namiri, M.K. Ashtiani, O. Mashinchian, M.M. Hasani-Sadrabadi,
M. Mahmoudi, N. Aghdami, et al., Engineering natural heart valves: possibilities and challenges, J. Tissue Eng. Regen. Med. (2016), http://dx.doi.org/
10.1002/term.2127.
[55] S.P. Hoerstrup, R. Sodian, S. Daebritz, J. Wang, E.A. Bacha, D.P. Martin, et al.,
Functional living trileaet heart valves grown in vitro, Circulation 102
(2000). Iii-44-Iii-9.
[56] K.J. Grande-Allen, J. Liao, The heterogeneous biomechanics and mechanobiology of the mitral valve: implications for tissue engineering, Curr. Cardiol.
Rep. 13 (2011) 113e120.
[57] H.C. Ott, T.S. Matthiesen, S.-K. Goh, L.D. Black, S.M. Kren, T.I. Netoff, et al.,
Perfusion-decellularized matrix: using nature's platform to engineer a bioarticial heart, Nat. Med. 14 (2008) 213e221.
[58] B. Zhang, Y. Xiao, A. Hsieh, N. Thavandiran, M. Radisic, Micro- and nanotechnology in cardiovascular tissue engineering, Nanotechnology 22 (2011)
494003.
[59] A. de Mel, G. Jell, M.M. Stevens, A.M. Seifalian, Biofunctionalization of biomaterials for accelerated in situ endothelialization: a review, Biomacromolecules 9 (2008) 2969e2979.
[60] J. Meng, G. Zhu, H. Xu, Effects of nanotopography for biomaterials on cell
behaviors, J. Biomed. Eng. 24 (2007) 685e689.
[61] K.R. Milner, C.A. Siedlecki, Fibroblast response is enhanced by poly (L-lactic
acid) nanotopography edge density and proximity, Int. J. Nanomedicine 2
(2007) 201.
[62] H. Jiang, G. Campbell, D. Boughner, W.-K. Wan, M. Quantz, Design and
manufacture of a polyvinyl alcohol (PVA) cryogel tri-leaet heart valve
prosthesis, Med. Eng. Phys. 26 (2004) 269e277.
[63] S. Van Vlierberghe, P. Dubruel, E. Schacht, Biopolymer-based hydrogels as
scaffolds for tissue engineering applications: a review, Biomacromolecules
12 (2011) 1387e1408.
[64] C. Bouten, P. Dankers, A. Driessen-Mol, S. Pedron, A. Brizard, F. Baaijens,
Substrates for cardiovascular tissue engineering, Adv. Drug Deliv. Rev. 63
(2011) 221e241.
[65] G.C. Rutledge, S.V. Fridrikh, Formation of bers by electrospinning, Adv. Drug
Deliv. Rev. 59 (2007) 1384e1391.
[66] D. Nisbet, J. Forsythe, W. Shen, D. Finkelstein, M. Horne, Review paper: a
review of the cellular response on electrospun nanobers for tissue engineering, J. Biomaterials Appl. (2008) 1e24.
[67] M.Z. Albanna, T.H. Bou-Akl, H.L. Walters III, H.W. Matthew, Improving the
mechanical properties of chitosan-based heart valve scaffolds using chitosan
bers, J. Mech. Behav. Biomed. Mater. 5 (2012) 171e180.
[68] P.S. Robinson, S.L. Johnson, M.C. Evans, V.H. Barocas, R.T. Tranquillo, Functional tissue-engineered valves from cell-remodeled brin with commissural
alignment of cell-produced collagen, Tissue Eng. Part A 14 (2008) 83e95.
[69] B. Guo, P.X. Ma, Synthetic biodegradable functional polymers for tissue engineering: a brief review, Sci. China Chem. 57 (2014) 490e500.
[70] C.E. Eckert, B.T. Mikulis, D. Gottlieb, D. Gerneke, I. LeGrice, R.F. Padera, et al.,
Three-dimensional quantitative micromorphology of pre-and post-implanted engineered heart valve tissues, Ann. Biomed. Eng. 39 (2011) 205e222.
[71] J.J. Stankus, J. Guan, K. Fujimoto, W.R. Wagner, Microintegrating smooth
muscle cells into a biodegradable, elastomeric ber matrix, Biomaterials 27
(2006) 735e744.
[72] D. Kalfa, A. Bel, A. Chen-Tournoux, A. Della Martina, P. Rochereau, C. Coz, et
al., A polydioxanone electrospun valved patch to replace the right ventricular
outow tract in a growing lamb model, Biomaterials 31 (2010) 4056e4063.
ffer, et al.,
[73] S. Hinderer, J. Seifert, M. Votteler, N. Shen, J. Rheinlaender, T.E. Scha
Engineering of a bio-functionalized hybrid off-the-shelf heart valve, Biomaterials 35 (2014) 2130e2139.
[74] N. Masoumi, N. Annabi, A. Assmann, B.L. Larson, J. Hjortnaes, N. Alemdar, et
al., Tri-layered elastomeric scaffolds for engineering heart valve leaets,
Biomaterials 35 (2014) 7774e7785.
[75] N.J. Amoroso, A. D'Amore, Y. Hong, C.P. Rivera, M.S. Sacks, W.R. Wagner,
Microstructural manipulation of electrospun scaffolds for specic bending
stiffness for heart valve tissue engineering, Acta Biomater. 8 (2012)
4268e4277.
[76] T. Courtney, M.S. Sacks, J. Stankus, J. Guan, W.R. Wagner, Design and analysis
of tissue engineering scaffolds that mimic soft tissue mechanical anisotropy,
Biomaterials 27 (2006) 3631e3638.
[77] Q. Chen, A. Bruyneel, C. Carr, J. Czernuszka, Bio-mechanical properties of
novel bi-layer collagen-elastin scaffolds for heart valve tissue engineering,
Procedia Eng. 59 (2013) 247e254.
[78] A. Ramamurthi, I. Vesely, Evaluation of the matrix-synthesis potential of
crosslinked hyaluronan gels for tissue engineering of aortic heart valves,
Biomaterials 26 (2005) 999e1010.
[79] J.R. Venugopal, M.P. Prabhakaran, S. Mukherjee, R. Ravichandran, K. Dan,
S. Ramakrishna, Biomaterial strategies for alleviation of myocardial infarction, J. R. Soc. Interface 9 (2012) 1e19.
[80] D. Wheatley, L. Raco, G. Bernacca, I. Sim, P. Belcher, J. Boyd, Polyurethane:
material for the next generation of heart valve prostheses? Eur. J. Cardiothoracic Surg. 17 (2000) 440e448.
[81] H. Ghanbari, H. Viatge, A.G. Kidane, G. Burriesci, M. Tavakoli, A.M. Seifalian,
Polymeric heart valves: new materials, emerging hopes, Trends Biotechnol.
27 (2009) 359e367.
[82] R.Y. Kannan, H.J. Salacinski, J. De Groot, I. Clatworthy, L. Bozec, M. Horton, et

290

[83]

[84]

[85]

[86]

[87]

[88]
[89]

[90]

[91]
[92]

[93]
[94]

[95]
[96]
[97]

[98]

[99]

[100]

[101]

[102]
[103]
[104]
[105]
[106]
[107]

[108]

[109]

[110]

[111]
[112]

A. Hasan et al. / Biomaterials 103 (2016) 278e292


al., The antithrombogenic potential of a polyhedral oligomeric silsesquioxane (POSS) nanocomposite, Biomacromolecules 7 (2006) 215e223.
A.G. Kidane, G. Burriesci, M. Edirisinghe, H. Ghanbari, P. Bonhoeffer,
A.M. Seifalian, A novel nanocomposite polymer for development of synthetic
heart valve leaets, Acta Biomater. 5 (2009) 2409e2417.
K.S. Masters, D.N. Shah, L.A. Leinwand, K.S. Anseth, Crosslinked hyaluronan
scaffolds as a biologically active carrier for valvular interstitial cells, Biomaterials 26 (2005) 2517e2525.
R. Sodian, S.P. Hoerstrup, J.S. Sperling, S.H. Daebritz, D.P. Martin, F.J. Schoen,
et al., Tissue engineering of heart valves: in vitro experiences, Ann. Thorac.
Surg. 70 (2000) 140e144.
E. Rabkin, S. Hoerstrup, M. Aikawa, J. Mayer Jr., F. Schoen, Evolution of cell
phenotype and extracellular matrix in tissue-engineered heart valves during
in-vitro maturation and in-vivo remodeling, J. Heart Valve Dis. 11 (2002)
308e314.
R. Sodian, S.P. Hoerstrup, J.S. Sperling, S. Daebritz, D.P. Martin, A.M. Moran, et
al., Early in vivo experience with tissue-engineered trileaet heart valves,
Circulation 102 (2000). Iii-22-Iii-9.
P.A. Gunatillake, R. Adhikari, Biodegradable synthetic polymers for tissue
engineering, Eur. Cells Mater. 5 (2003) 1e16.
S. Suri, C.E. Schmidt, Photopatterned collagenehyaluronic acid interpenetrating polymer network hydrogels, Acta Biomater. 5 (2009)
2385e2397.
B. Duan, E. Kapetanovic, L. Hockaday, J. Butcher, Three-dimensional printed
trileaet valve conduits using biological hydrogels and human valve interstitial cells, Acta Biomater. 10 (2014) 1836e1846.
I. Liao, S. Chen, J.B. Liu, K.W. Leong, Sustained viral gene delivery through
core-shell bers, J. Control. Release 139 (2009) 48e55.
S. Stachelek, C. Song, I. Alferiev, S. Defelice, X. Cui, J. Connolly, et al., Localized
gene delivery using antibody tethered adenovirus from polyurethane heart
valve cusps and intra-aortic implants, Gene Ther. 11 (2004) 15e24.
M.P. Cuchiara, A.C. Allen, T.M. Chen, J.S. Miller, J.L. West, Multilayer microuidic PEGDA hydrogels, Biomaterials 31 (2010) 5491e5497.
C.C. Chang, E.D. Boland, S.K. Williams, J.B. Hoying, Direct-write bioprinting
three-dimensional biohybrid systems for future regenerative therapies,
J. Biomed. Mater. Res. Part B Appl. Biomaterials 98 (2011) 160e170.
B. Derby, Printing and prototyping of tissues and scaffolds, Science 338
(2012) 921e926.
S. Tasoglu, U. Demirci, Bioprinting for stem cell research, Trends Biotechnol.
31 (2013) 10e19.
B. Duan, L.A. Hockaday, K.H. Kang, J.T. Butcher, 3D bioprinting of heterogeneous aortic valve conduits with alginate/gelatin hydrogels, J. Biomed. Mater.
Res. Part A 101 (2013) 1255e1264.
L. Hockaday, K. Kang, N. Colangelo, P. Cheung, B. Duan, E. Malone, et al.,
Rapid 3D printing of anatomically accurate and mechanically heterogeneous
aortic valve hydrogel scaffolds, Biofabrication 4 (2012) 035005.
C. Lueders, B. Jastram, R. Hetzer, H. Schwandt, Rapid manufacturing techniques for the tissue engineering of human heart valves, Eur. J. Cardiothoracic Surg. (2014) ezt510.
T. Xu, K.W. Binder, M.Z. Albanna, D. Dice, W. Zhao, J.J. Yoo, et al., Hybrid
printing of mechanically and biologically improved constructs for cartilage
tissue engineering applications, Biofabrication 5 (2013) 015001.
D. Kami, S. Takeda, Y. Itakura, S. Gojo, M. Watanabe, M. Toyoda, Application
of magnetic nanoparticles to gene delivery, Int. J. Mol. Sci. 12 (2011)
3705e3722.
L.C. Heller, K. Ugen, R. Heller, Electroporation for targeted gene transfer,
Expert Opin. Drug Deliv. 2 (2005) 255e268.
N. Nayerossadat, T. Maedeh, P.A. Ali, Viral and nonviral delivery systems for
gene delivery, Adv. Biomed. Res. 1 (2012) 27.
D.P. Katare, V. Aeri, Progress in gene therapy: a review, IJTPR 1 (2010)
33e41.
C.E. Thomas, A. Ehrhardt, M.A. Kay, Progress and problems with the use of
viral vectors for gene therapy, Nat. Rev. Genet. 4 (2003) 346e358.
A. Hirko, F. Tang, J.A. Hughes, Cationic lipid vectors for plasmid DNA delivery,
Curr. Med. Chem. 10 (2003) 1185e1193.
F. Bof, A. Bonincontro, F. Bordi, E. Bultrini, C. Cametti, A. Congiu-Castellano,
et al., Two-step mechanism in cationic lipoplex formation as observed by
dynamic light scattering, dielectric relaxation and circular dichroism
methods, Phys. Chem. Chem. Phys. 4 (2002) 2708e2713.
M. Bertschinger, G. Backliwal, A. Schertenleib, M. Jordan, D.L. Hacker,
F.M. Wurm, Disassembly of polyethylenimine-DNA particles in vitro: implications for polyethylenimine-mediated DNA delivery, J. Control. Release
116 (2006) 96e104.
M.L. Forrest, J.T. Koerber, D.W. Pack, A degradable polyethylenimine derivative with low toxicity for highly efcient gene delivery, Bioconjugate Chem.
14 (2003) 934e940.
Q. Peng, Z. Zhong, R. Zhuo, Disulde cross-linked polyethylenimines (PEI)
prepared via thiolation of low molecular weight PEI as highly efcient gene
vectors, Bioconjugate Chem. 19 (2008) 499e506.
J. Dobson, Gene therapy progress and prospects: magnetic nanoparticlebased gene delivery, Gene Ther. 13 (2006) 283e287.
S.I. Jenkins, M.R. Pickard, N. Granger, D.M. Chari, Magnetic nanoparticlemediated gene transfer to oligodendrocyte precursor cell transplant populations is enhanced by magnetofection strategies, ACS Nano 5 (2011)
6527e6538.

[113] C. Plank, J. Rosenecker, Magnetofection: the use of magnetic nanoparticles


for nucleic acid delivery, Cold Spring Harb. Protoc. (2009), http://dx.doi.org/
10.1101/pdb.prot5230.
[114] F. Scherer, M. Anton, U. Schillinger, J. Henke, C. Bergemann, A. Kruger, et al.,
Magnetofection: enhancing and targeting gene delivery by magnetic force
in vitro and in vivo, Gene Ther. 9 (2002) 102e109.
[115] M. Subramanian, J. Lim, J. Dobson, Enhanced nanomagnetic gene transfection
of human prenatal cardiac progenitor cells and adult cardiomyocytes, PLoS
One 8 (2013) e69812.
[116] Z. Yin, N. Liu, M. Ma, L. Wang, Y. Hao, X. Zhang, A novel EGFR-targeted gene
delivery system based on complexes self-assembled by EGF, DNA, and
activated PAMAM dendrimers, Int. J. Nanomedicine 7 (2012) 4625.
s, Trans[117] S. Somani, D.R. Blatchford, O. Millington, M.L. Stevenson, C. Dufe
ferrin-bearing polypropylenimine dendrimer for targeted gene delivery to
the brain, J. Control. Release 188 (2014) 78e86.
[118] A. Kumar, V.K. Yellepeddi, G.E. Davies, K.B. Strychar, S. Palakurthi, Enhanced
gene transfection efciency by polyamidoamine (PAMAM) dendrimers
modied with ornithine residues, Int. J. Pharm. 392 (2010) 294e303.
[119] P. Wang, X.-H. Zhao, Z.-Y. Wang, M. Meng, X. Li, Q. Ning, Generation 4 polyamidoamine dendrimers is a novel candidate of nano-carrier for gene delivery agents in breast cancer treatment, Cancer Lett. 298 (2010) 34e49.
[120] J.H. Park, J.-S. Park, J.S. Choi, Basic amino acid-conjugated polyamidoamine
dendrimers with enhanced gene transfection efciency, Macromol. Res. 22
(2014) 500e508.
[121] M. Yu, X. Jie, L. Xu, C. Chen, W. Shen, Y. Cao, et al., Recent advances in
dendrimer research for cardiovascular diseases, Biomacromolecules 16
(2015) 2588e2598.
[122] D. Xitong, Z. Xiaorong, Targeted therapeutic delivery using engineered
exosomes and its applications in cardiovascular diseases, Gene 575 (2016)
377e384.
[123] Y. Lee, S.E. Andaloussi, M.J. Wood, Exosomes and microvesicles: extracellular
vesicles for genetic information transfer and gene therapy, Hum. Mol. Genet.
21 (2012) R125eR134.
ller,
[124] O. Mykhaylyk, D. Vlaskou, N. Tresilwised, P. Pithayanukul, W. Mo
C. Plank, Magnetic nanoparticle formulations for DNA and siRNA delivery,
J. Magn. Magn. Mater. 311 (2007) 275e281.
[125] C. Sapet, N. Laurent, A. de Chevigny, L. Le Gourrierec, E. Bertosio, O. Zelphati,
et al., High transfection efciency of neural stem cells with magnetofection,
Biotechniques 50 (2011) 187e189.
[126] E.R. Figueroa, A.Y. Lin, J. Yan, L. Luo, A.E. Foster, R.A. Drezek, Optimization of
PAMAM-gold nanoparticle conjugation for gene therapy, Biomaterials 35
(2014) 1725e1734.
[127] Q. Zhang, S. Chen, R.-X. Zhuo, X.-Z. Zhang, S.-X. Cheng, Self-assembled terplexes for targeted gene delivery with improved transfection, Bioconjugate
Chem. 21 (2010) 2086e2092.
[128] D. Cui, P. Huang, C. Zhang, C.S. Ozkan, B. Pan, P. Xu, Dendrimer-modied gold
nanorods as efcient controlled gene delivery system under near-infrared
light irradiation, J. Control. Release 152 (2011) e137ee139.
[129] J. Li, L. Chen, N. Liu, S. Li, Y. Hao, X. Zhang, EGF-coated nano-dendriplexes for
tumor-targeted nucleic acid delivery in vivo, Drug Deliv. (2015) 1e8.
[130] A.J. O'Loughlin, C.A. Wofndale, M.J. Wood, Exosomes and the emerging eld
of exosome-based gene therapy, Curr. Gene Ther. 12 (2012) 262e274.
[131] X. Sun, A. Gutierrez, M.J. Yacaman, X. Dong, S. Jin, Investigations on magnetic
properties and structure for carbon encapsulated nanoparticles of Fe, Co, Ni,
Mater. Sci. Eng. A 286 (2000) 157e160.
[132] A. Tomitaka, H. Kobayashi, T. Yamada, M. Jeun, S. Bae, Y. Takemura,
Magnetization and self-heating temperature of NiFe2O4 nanoparticles
measured by applying ac magnetic eld, J. Phys. Conf. Ser. IOP Publ. (2010)
122010.
[133] W.-S. Cho, R. Dufn, C.A. Poland, A. Duschl, G.J. Oostingh, W. MacNee, et al.,
Differential pro-inammatory effects of metal oxide nanoparticles and their
soluble ions in vitro and in vivo; zinc and copper nanoparticles, but not their
ions, recruit eosinophils to the lungs, Nanotoxicology 6 (2012) 22e35.
[134] S. George, T. Xia, R. Rallo, Y. Zhao, Z. Ji, S. Lin, et al., Use of a high-throughput
screening approach coupled with in vivo zebrash embryo screening to
develop hazard ranking for engineered nanomaterials, ACS Nano 5 (2011)
1805e1817.
[135] J. Giri, P. Pradhan, V. Somani, H. Chelawat, S. Chhatre, R. Banerjee, et al.,
Synthesis and characterizations of water-based ferrouids of substituted
ferrites for biomedical applications, J. Magn. Magn. Mater. 320 (2008)
724e730.
[136] H.L. Karlsson, P. Cronholm, J. Gustafsson, L. Moller, Copper oxide nanoparticles are highly toxic: a comparison between metal oxide nanoparticles
and carbon nanotubes, Chem. Res. Toxicol. 21 (2008) 1726e1732.
[137] H. Basti, L.B. Tahar, L. Smiri, F. Herbst, M.-J. Vaulay, F. Chau, et al., Catechol
derivatives-coated Fe3O4 and g-Fe2O3 nanoparticles as potential MRI
contrast agents, J. Colloid Interface Sci. 341 (2010) 248e254.
[138] L. Gamarra, E. Amaro, S. Alves, D. Soga, W.M. Pontuschka, J. Mamani, et al.,
Characterization of the biocompatible magnetic colloid on the basis of Fe3O4
nanoparticles coated with dextran, used as contrast agent in magnetic
resonance imaging, J. Nanosci. Nanotechnol. 10 (2010) 4145e4153.
nager, O. Cle
ment, G. Barratt, C. Grabielle[139] M.-S. Martina, J.-P. Fortin, C. Me
Madelmont, et al., Generation of superparamagnetic liposomes revealed as
highly efcient MRI contrast agents for in vivo imaging, J. Am. Chem. Soc.
127 (2005) 10676e10685.

A. Hasan et al. / Biomaterials 103 (2016) 278e292


[140] K. Turcheniuk, A.V. Tarasevych, V.P. Kukhar, R. Boukherroub, S. Szunerits,
Recent advances in surface chemistry strategies for the fabrication of functional iron oxide based magnetic nanoparticles, Nanoscale 5 (2013)
10729e10752.
[141] N. Nitin, L. LaConte, O. Zurkiya, X. Hu, G. Bao, Functionalization and peptidebased delivery of magnetic nanoparticles as an intracellular MRI contrast
agent, JBIC e J. Biol. Inorg. Chem. 9 (2004) 706e712.
[142] C.J. Mertz, M.D. Kaminski, Y. Xie, M.R. Finck, S. Guy, A.J. Rosengart, In vitro
studies of functionalized magnetic nanospheres for selective removal of a
simulant biotoxin, J. Magn. Magn. Mater. 293 (2005) 572e577.
[143] M. Mikhaylova, Y. Jo, D.-K. Kim, N. Bobrysheva, Y. Andersson, T. Eriksson, et
al., The effect of biocompatible coating layers on magnetic properties of
superparamagnetic iron oxide nanoparticles, Hyperne Interact. 156 (2004)
257e263.
[144] M. Arsianti, M. Lim, S.N. Lou, I.Y. Goon, C.P. Marquis, R. Amal, Bi-functional
gold-coated magnetite composites with improved biocompatibility, J. Colloid
Interface Sci. 354 (2011) 536e545.
[145] C.C. Berry, S. Wells, S. Charles, A.S. Curtis, Dextran and albumin derivatised
iron oxide nanoparticles: inuence on broblasts in vitro, Biomaterials 24
(2003) 4551e4557.
[146] A. Ito, K. Ino, T. Kobayashi, H. Honda, The effect of RGD peptide-conjugated
magnetite cationic liposomes on cell growth and cell sheet harvesting,
Biomaterials 26 (2005) 6185e6193.
s, Glyconanoparticles: types, synthesis and applications in
[147] M. Jess, S. Penade
glycoscience, biomedicine and material science, Biochim. Biophys. Acta
(BBA) Gen. Subj. 1760 (2006) 636e651.
[148] M.A. McDonald, K.L. Watkin, Investigations into the physicochemical properties of dextran small particulate gadolinium oxide nanoparticles, Acad.
Radiol. 13 (2006) 421e427.
[149] H. Zhang, T. Xia, H. Meng, M. Xue, S. George, Z. Ji, et al., Differential
expression of syndecan-1 mediates cationic nanoparticle toxicity in undifferentiated versus differentiated normal human bronchial epithelial cells,
ACS Nano 5 (2011) 2756e2769.
[150] C. Schweiger, C. Pietzonka, J. Heverhagen, T. Kissel, Novel magnetic iron
oxide nanoparticles coated with poly (ethylene imine)-g-poly (ethylene
glycol) for potential biomedical application: synthesis, stability, cytotoxicity
and MR imaging, Int. J. Pharm. 408 (2011) 130e137.
[151] K.H. Zuo, D.L. Jiang, J. xian Zhang, Q.L. Lin, Forming nanometer TiO 2 sheets
by nonaqueous tape casting, Ceram. Int. 33 (2007) 477e481.
[152] F.M. Kievit, O. Veiseh, N. Bhattarai, C. Fang, J.W. Gunn, D. Lee, et al.,
PEIePEGechitosan-copolymer-coated iron oxide nanoparticles for safe gene
delivery: synthesis, complexation, and transfection, Adv. Funct. Mater. 19
(2009) 2244e2251.
[153] Y. Shi, L. Zhou, R. Wang, Y. Pang, W. Xiao, H. Li, et al., In situ preparation of
magnetic nonviral gene vectors and magnetofection in vitro, Nanotechnology 21 (2010) 115103.
[154] R. Namgung, K. Singha, M.K. Yu, S. Jon, Y.S. Kim, Y. Ahn, et al., Hybrid
superparamagnetic iron oxide nanoparticle-branched polyethylenimine
magnetoplexes for gene transfection of vascular endothelial cells, Biomaterials 31 (2010) 4204e4213.
[155] M. Arsianti, M. Lim, C.P. Marquis, R. Amal, Polyethylenimine based magnetic
iron-oxide vector: the effect of vector component assembly on cellular entry
mechanism, intracellular localization, and cellular viability, Biomacromolecules 11 (2010) 2521e2531.
rez, P.L. Sa
nchez, S.G. de Cruz, E. Villacorta, I.F. Palacios,
[156] M.A. Bratos-Pe
ndez-Ferna
ndez, et al., Association between self-replicating calciJ.M. Ferna
fying nanoparticles and aortic stenosis: a possible link to valve calcication,
Eur. Heart J. 29 (2008) 371e376.
[157] H. Ghanbari, A.G. Kidane, G. Burriesci, B. Ramesh, A. Darbyshire,
A.M. Seifalian, The anti-calcication potential of a silsesquioxane nanocomposite polymer under in vitro conditions: potential material for synthetic leaet heart valve, Acta Biomater. 6 (2010) 4249e4260.
[158] B. Rahmani, S. Tzamtzis, H. Ghanbari, G. Burriesci, A.M. Seifalian,
Manufacturing and hydrodynamic assessment of a novel aortic valve made
of a new nanocomposite polymer, J. Biomech. 45 (2012) 1205e1211.
[159] F.D. Fortuin, P. Vale, D.W. Losordo, J. Symes, G.A. DeLaria, J.J. Tyner, et al.,
One-year follow-up of direct myocardial gene transfer of vascular endothelial growth factor-2 using naked plasmid deoxyribonucleic acid by way of
thoracotomy in no-option patients, Am. J. Cardiol. 92 (2003) 436e439.
[160] H.E. Hoand, D. Nagy, J.-J. Liu, K. Spratt, Y.-L. Lee, O. Danos, et al., In vivo gene
transfer by intravenous administration of stable cationic lipid/DNA complex,
Pharm. Res. 14 (1997) 742e749.
[161] J.W. Yockman, A. Kastenmeier, H.M. Erickson, J.G. Brumbach, M.G. Whitten,
A. Albanil, et al., Novel polymer carriers and gene constructs for treatment of
myocardial ischemia and infarction, J. Control. Release 132 (2008) 260e266.
glon, L.K. von Segesser,
[162] S. Fleury, E. Simeoni, C. Zuppinger, N. De
L. Kappenberger, et al., Multiply attenuated, self-inactivating lentiviral vectors efciently deliver and express genes for extended periods of time in
adult rat cardiomyocytes in vivo, Circulation 107 (2003) 2375e2382.
[163] J. Zhao, G.J. Pettigrew, J. Thomas, J.I. Vandenberg, L. Delriviere, E.M. Bolton, et
al., Lentiviral vectors for delivery of genes into neonatal and adult ventricular
cardiac myocytes in vitro and in vivo, Basic Res. Cardiol. 97 (2002) 348e358.
[164] K. Niwano, M. Arai, N. Koitabashi, A. Watanabe, Y. Ikeda, H. Miyoshi, et al.,
Lentiviral vectoremediated SERCA2 gene transfer protects against heart
failure and left ventricular remodeling after myocardial infarction in rats,

291

Mol. Ther. 16 (2008) 1026e1032.


G. Tang, et al., Gene delivery to
[165] W. Li, C. Nesselmann, Z. Zhou, L.-L. Ong, F. Ori,
the heart by magnetic nanobeads, J. Magn. Magn. Mater. 311 (2007)
336e341.
[166] W. Li, N. Ma, L.L. Ong, A. Kaminski, C. Skrabal, M. Ugurlucan, et al., Enhanced
thoracic gene delivery by magnetic nanobead-mediated vector, J. Gene Med.
10 (2008) 897e909.
[167] Y. Zhang, W. Li, L. Ou, W. Wang, E. Delyagina, C. Lux, et al., Targeted delivery
of human VEGF gene via complexes of magnetic nanoparticle-adenoviral
vectors enhanced cardiac regeneration, PLos One 7 (2012) e39490.
[168] R. Lanza, R. Langer, J.P. Vacanti, Principles of Tissue Engineering, Academic
Press, 2011.
[169] T. Ota, Y. Sawa, S. Iwai, T. Kitajima, Y. Ueda, C. Coppin, et al., Fibronectinhepatocyte growth factor enhances reendothelialization in tissueengineered heart valve, Ann. Thorac. Surg. 80 (2005) 1794e1801.
[170] Stelt B. Tissue engineering of aortic valves.
[171] S. Cebotari, H. Mertsching, K. Kallenbach, S. Kostin, O. Repin, A. Batrinac, et
al., Construction of autologous human heart valves based on an acellular
allograft matrix, Circulation 106 (2002). I-63-I-8.
[172] R. Morishita, M. Aoki, Y. Yo, T. Ogihara, Hepatocyte growth factor as cardiovascular hormone: role of HGF in the pathogenesis of cardiovascular
disease, Endocr. J. 49 (2002) 273e284.
[173] E.J. Armstrong, J. Bischoff, Heart valve development endothelial cell signaling
and differentiation, Circulat. Res. 95 (2004) 459e470.
[174] D. Beis, T. Bartman, S.-W. Jin, I.C. Scott, L.A. D'Amico, E.A. Ober, et al., Genetic
and cellular analyses of zebrash atrioventricular cushion and valve development, Development 132 (2005) 4193e4204.
[175] C. Urbich, A. Aicher, C. Heeschen, E. Dernbach, W.K. Hofmann, A.M. Zeiher, et
al., Soluble factors released by endothelial progenitor cells promote migration of endothelial cells and cardiac resident progenitor cells, J. Mol. Cell.
Cardiol. 39 (2005) 733e742.
[176] A. Gandaglia, A. Bagno, F. Naso, M. Spina, G. Gerosa, Cells, scaffolds and
bioreactors for tissue-engineered heart valves: a journey from basic concepts
to contemporary developmental innovations, Eur. J. Cardio-thoracic Surg. 39
(2011) 523e531.
[177] W.L. Grayson, F. Zhao, B. Bunnell, T. Ma, Hypoxia enhances proliferation and
tissue formation of human mesenchymal stem cells, Biochem. Biophys. Res.
Commun. 358 (2007) 948e953.
[178] A.J. Appleton, C.T.G. Appleton, D.R. Boughner, K.A. Rogers, Vascular smooth
muscle cells as a valvular interstitial cell surrogate in heart valve tissue
engineering, Tissue Eng. Part A 15 (2009) 3889e3897.
[179] B.K. Mann, R.H. Schmedlen, J.L. West, Tethered-TGF-b increases extracellular
matrix production of vascular smooth muscle cells, Biomaterials 22 (2001)
439e444.
[180] J.P. Stegemann, R.M. Nerem, Phenotype modulation in vascular tissue engineering using biochemical and mechanical stimulation, Ann. Biomed. Eng. 31
(2003) 391e402.
[181] S.-D. Huang, X.-H. Liu, C.-G. Bai, F.-L. Lu, Y. Yuan, D.-J. Gong, et al., Synergistic
effect of bronectin and hepatocyte growth factor on stable cell-matrix
adhesion, re-endothelialization, and reconstitution in developing tissueengineered heart valves, Heart Vessels 22 (2007) 116e122.
[182] N. Dong, Y. Qiu, J. Shi, Application of transforming growth factor-beta (1) on
construction of tissue engineering heart valves, Exp. In Vitro 87 (2007)
1622e1626.
[183] S.P. Hoerstrup, R. Sodian, J.S. Sperling, J.P. Vacanti, J.E. Mayer Jr., New pulsatile bioreactor for in vitro formation of tissue engineered heart valves,
Tissue Eng. 6 (2000) 75e79.
[184] R. Sodian, S.P. Hoerstrup, J.S. Sperling, S.H. Daebritz, D.P. Martin, F.J. Schoen,
et al., Tissue engineering of heart valves: in vitro experiences, Ann. Thorac.
Surg. 70 (2000) 140e144.
[185] M.B. Chen, S. Srigunapalan, A.R. Wheeler, C.A. Simmons, A 3D microuidic
platform incorporating methacrylated gelatin hydrogels to study physiological cardiovascular cellecell interactions, Lab Chip 13 (2013) 2591e2598.
[186] A.A. Bakhaty, M.R.K. Mofrad, Coupled simulation of heart valves: application
to clinical practice, Ann. Rev. Biomed. Eng. 43 (7) (2015) 1626e1639.
[187] W. Sun, C. Martin, T. Pham, Computational modeling of cardiac valve function and intervention, Annu. Rev. Biomed. Eng. 16 (2014) 53e76.
[188] E.J. Weinberg, D. Shahmirzadi, M.R. Mofrad, On the multiscale modeling of
heart valve biomechanics in health and disease, Biomech. Model. Mechanobiol. 9 (2010) 373e387.
[189] R. Voronov, S. VanGordon, V.I. Sikavitsas, D.V. Papavassiliou, Computational
modeling of ow-induced shear stresses within 3D salt-leached porous
scaffolds imaged via micro-CT, J. Biomech. 43 (2010) 1279e1286.
[190] H.S. Tuan, D.W. Hutmacher, Application of micro CT and computation
modeling in bone tissue engineering, Computer-Aided Des. 37 (2005)
1151e1161.
[191] P. Sucosky, D.F. Osorio, J.B. Brown, G.P. Neitzel, Fluid mechanics of a spinnerask bioreactor, Biotechnol. Bioeng. 85 (2004) 34e46.
[192] M.L. Azoitei, B.E. Correia, Y.-E.A. Ban, C. Carrico, O. Kalyuzhniy, L. Chen, et al.,
Computation-guided backbone grafting of a discontinuous motif onto a
protein scaffold, Science 334 (2011) 373e376.
[193] C.C. Doumanidis, Nanomanufacturing of random branching material architectures, Microelectron. Eng. 86 (2009) 467e478.
[194] O.C. Zienkiewicz, R.L. Taylor, The Finite Element Method for Solid and
Structural Mechanics, Butterworth-Heinemann, 2005.

292

A. Hasan et al. / Biomaterials 103 (2016) 278e292

[195] C. Hirsch, Numerical Computation of Internal and External Flows: the Fundamentals of Computational Fluid Dynamics: the Fundamentals of Computational Fluid Dynamics, Butterworth-Heinemann, 2007.
[196] K. Burke, J. Werschnik, E. Gross, Time-dependent density functional theory:
past, present, and future, J. Chem. Phys. 123 (2005) 062206.
[197] S.C. Vigmostad, H.S. Udaykumar, J. Lu, K.B. Chandran, Fluid-structure interaction methods in biological ows with special emphasis on heart valve
dynamics, Int. J. Numer. Methods Biomed. Eng. 26 (2010) 435e470.
[198] R. Haj-Ali, L.P. Dasi, H.S. Kim, J. Choi, H.W. Leo, A.P. Yoganathan, Structural
simulations of prosthetic tri-leaet aortic heart valves, J. Biomech. 41 (2008)
1510e1519.
[199] G. Querzoli, S. Fortini, A. Cenedese, Effect of the prosthetic mitral valve on
vortex dynamics and turbulence of the left ventricular ow, Phys. Fluids 22
(2010) 041901.
[200] W. Sun, K. Li, E. Sirois, Simulated elliptical bioprosthetic valve deformation:
implications for asymmetric transcatheter valve deployment, J. Biomech. 43
(2010) 3085e3090.
[201] M.D. de Tullio, G. Pascazio, L. Weltert, R. De Paulis, R. Verzicco, Evaluation of
prosthetic-valved devices by means of numerical simulations, Philos. Trans.
Ser. A Math. Phys. Eng. Sci. 369 (2011) 2502e2509.
[202] T.B. Le, F. Sotiropoulos, Fluid-structure interaction of an aortic heart valve
prosthesis driven by an animated anatomic left ventricle, J. Comput. Phys.
244 (2013) 41e62.
[203] C. Capelli, G. Biglino, L. Petrini, F. Migliavacca, D. Cosentino, P. Bonhoeffer, et
al., Finite element strategies to satisfy clinical and engineering requirements
in the eld of percutaneous valves, Ann. Biomed. Eng. 40 (2012) 2663e2673.
[204] C. Guivier-Curien, V. Deplano, E. Bertrand, Validation of a numerical 3-D
uid-structure interaction model for a prosthetic valve based on experimental PIV measurements, Med. Eng. Phys. 31 (2009) 986e993.
[205] C. Martin, W. Sun, Simulation of long-term fatigue damage in bioprosthetic
heart valves: effects of leaet and stent elastic properties, Biomech. Model.
Mechanobiol. 13 (2014) 759e770.
[206] F. Sotiropoulos, I. Borazjani, A review of state-of-the-art numerical methods
for simulating ow through mechanical heart valves, Med. Biol. Eng. Comput. 47 (2009) 245e256.
[207] D. Bluestein, Y. Li, I. Krukenkamp, Free emboli formation in the wake of bileaet mechanical heart valves and the effects of implantation techniques,
J. Biomech. 35 (2002) 1533e1540.
[208] D. Bluestein, E. Rambod, M. Gharib, Vortex shedding as a mechanism for free
emboli formation in mechanical heart valves, J. Biomech. Eng. 122 (2000)
125e134.
[209] M.S. Sacks, A. Mirnaja, W. Sun, P. Schmidt, Bioprosthetic Heart Valve Heterograft Biomaterials: Structure, Mechanical Behavior and Computational
Simulation, 2006.
[210] Y. Alemu, D. Bluestein, Flow-induced platelet activation and damage accumulation in a mechanical heart valve: numerical studies, Artif. Organs 31
(2007) 677e688.
[211] U. Morbiducci, R. Ponzini, M. Nobili, D. Massai, F.M. Montevecchi,
D. Bluestein, et al., Blood damage safety of prosthetic heart valves. Shearinduced platelet activation and local ow dynamics: a uid-structure
interaction approach, J. Biomech. 42 (2009) 1952e1960.
[212] S. Shahriari, H. Maleki, I. Hassan, L. Kadem, Evaluation of shear stress accumulation on blood components in normal and dysfunctional bileaet mechanical heart valves using smoothed particle hydrodynamics, J. Biomech. 45
(2012) 2637e2644.
[213] M.S. Sacks, Incorporation of experimentally-derived ber orientation into a
structural constitutive model for planar collagenous tissues, J. Biomech. Eng.
125 (2003) 280e287.
[214] J.L. Tan, J. Tien, D.M. Pirone, D.S. Gray, K. Bhadriraju, C.S. Chen, Cells lying on a
bed of microneedles: an approach to isolate mechanical force, Proc. Natl.
Acad. Sci. 100 (2003) 1484e1489.
[215] D. Vader, A. Kabla, D. Weitz, L. Mahadevan, Strain-induced alignment in
collagen gels, PLoS One 4 (2009) e5902.
[216] Z. Yang, J.-S. Lin, J. Chen, J.H. Wang, Determining substrate displacement and
cell traction eldsda new approach, J. Theor. Biol. 242 (2006) 607e616.
[217] J. Friedrichs, A. Taubenberger, C.M. Franz, D.J. Muller, Cellular remodelling of
individual collagen brils visualized by time-lapse AFM, J. Mol. Biol. 372
(2007) 594e607.
[218] N.J. Driessen, M.A. Cox, C.V. Bouten, F.P. Baaijens, Remodelling of the angular
collagen ber distribution in cardiovascular tissues, Biomech. Model.

Mechanobiol. 7 (2008) 93e103.


[219] N.J.B. Driessen, Computational analyses of mechanically induced collagen
ber remodeling in the aortic heart valve, J. Biomech. Eng. 125 (2003) 549.
[220] R. Amini, C.A. Voycheck, R.E. Debski, A method for predicting collagen ber
realignment in non-planar tissue surfaces as applied to glenohumeral
capsule during clinically relevant deformation, J. Biomech. Eng. 136 (2014)
031003.
[221] A.L. Soares, C.W. Oomens, F.P. Baaijens, A computational model to describe
the collagen orientation in statically cultured engineered tissues, Comput.
Methods Biomech. Biomed. Eng. 17 (2014) 251e262.
[222] J.S. Soares, J. Sheriff, D. Bluestein, A novel mathematical model of activation
and sensitization of platelets subjected to dynamic stress histories, Biomech.
Model. Mechanobiol. 12 (2013) 1127e1141.
[223] X.-J. Luo, T. Stylianopoulos, V.H. Barocas, M.S. Shephard, Multiscale computation for bioarticial soft tissues with complex geometries, Eng. Comput. 25
(2009) 87e95.
[224] V. Kouznetsova, W. Brekelmans, F. Baaijens, An approach to micro-macro
modeling of heterogeneous materials, Comput. Mech. 27 (2001) 37e48.
[225] H.-Y.S. Huang, J. Liao, M.S. Sacks, In-situ deformation of the aortic valve
interstitial cell nucleus under diastolic loading, J. Biomech. Eng. 129 (2007)
880e889.
[226] S. Huang, H.-Y.S. Huang, Ieee. Virtual experiments of heart valve tissues, in:
2012 Annual International Conference of the Ieee Engineering in Medicine
and Biology Society, 2012, pp. 6645e6648.
[227] V. Prot, B. Skallerud, Nonlinear solid nite element analysis of mitral valves
with heterogeneous leaet layers, Comput. Mech. 43 (2009) 353e368.
[228] B. Skallerud, V. Prot, I.S. Nordrum, Modeling active muscle contraction in
mitral valve leaets during systole: a rst approach, Biomech. Model.
Mechanobiol. 10 (2011) 11e26.
[229] G. Argento, M. Simonet, C.W. Oomens, F.P. Baaijens, Multi-scale mechanical
characterization of scaffolds for heart valve tissue engineering, J. Biomech. 45
(2012) 2893e2898.
[230] R. Fan, A.S. Bayoumi, P. Chen, C.M. Hobson, W.R. Wagner, J.E. Mayer Jr., et al.,
Optimal elastomeric scaffold leaet shape for pulmonary heart valve leaet
replacement, J. Biomech. 46 (2013) 662e669.
[231] L. Wang, S. Korossis, E. Ingham, J. Fisher, Z. Jin, Computational simulation of
oxygen diffusion in aortic valve leaet for tissue engineering applications,
J. Heart Valve Dis. 17 (2008) 700e709.
[232] C.K. Aidun, J.R. Clausen, Lattice-Boltzmann method for complex ows, Annu.
Rev. Fluid Mech. 42 (2010) 439e472.
[233] K. Madi, G. Tozzi, Q.H. Zhang, J. Tong, A. Cossey, A. Au, et al., Computation of
full-eld displacements in a scaffold implant using digital volume correlation
and nite element analysis, Med. Eng. Phys. 35 (2013) 1298e1312.
[234] K.S. Chan, W. Liang, W.L. Francis, D.P. Nicolella, A multiscale modeling
approach to scaffold design and property prediction, J. Mech. Behav. Biomed.
Mater. 3 (2010) 584e593.
[235] A.K. Miri, H.K. Heris, L. Mongeau, F. Javid, Nanoscale viscoelasticity of
extracellular matrix proteins in soft tissues: a multiscale approach, J. Mech.
Behav. Biomed. Mater. 30 (2014) 196e204.
[236] H.T. Nia, L. Han, Y. Li, C. Ortiz, A. Grodzinsky, Poroelasticity of cartilage at the
nanoscale, Biophys. J. 101 (2011) 2304e2313.
[237] J.F. Wenk, P. Papadopoulos, T.I. Zohdi, Numerical modeling of stress in stenotic arteries with microcalcications: a micromechanical approximation,
J. Biomech. Eng. 132 (2010) 091011.
[238] G.A. Ateshian, T. Ricken, Multigenerational interstitial growth of biological
tissues, Biomech. Model. Mechanobiol. 9 (2010) 689e702.
[239] D. Lasseux, A. Ahmadi, X. Cleis, J. Garnier, A macroscopic model for species
transport during in vitro tissue growth obtained by the volume averaging
method, Chem. Eng. Sci. 59 (2004) 1949e1964.
[240] D.E. Ingber, The mechanochemical basis of cell and tissue regulation, Mech.
Chem. Biosyst. 1 (2004) 53e68.
[241] D.E. Discher, P. Janmey, Y.L. Wang, Tissue cells feel and respond to the
stiffness of their substrate, Science 310 (2005) 1139e1143.
[242] X. Zeng, S. Li, Multiscale modeling and simulation of soft adhesion and
contact of stem cells, J. Mech. Behav. Biomed. Mater. 4 (2011) 180e189.
[243] J. Zhu, R.E. Marchant, Design properties of hydrogel tissue-engineering
scaffolds, Expert Rev. Med. Devices 8 (2011) 607e626.
[244] S. Jana, A. Lerman, Bioprinting a cardiac valve, Biotechnol. Adv. 33 (2015)
1503e1521.

Potrebbero piacerti anche