Vous êtes sur la page 1sur 19

Biomaterials 27 (2006) 34133431

Review
Biodegradable and bioactive porous polymer/inorganic composite
scaffolds for bone tissue engineering
K. Rezwan
a
, Q.Z. Chen
a
, J.J. Blaker
a
, Aldo Roberto Boccaccini
a,b,
a
Department of Materials, Imperial College London, Prince Consort Road, London SW7 2BP, UK
b
Centre for Tissue Engineering and Regenerative Medicine, Imperial College London, London SW7 2AZ, UK
Received 6 December 2005; accepted 31 January 2006
Available online 28 February 2006
Abstract
Biodegradable polymers and bioactive ceramics are being combined in a variety of composite materials for tissue engineering scaffolds.
Materials and fabrication routes for three-dimensional (3D) scaffolds with interconnected high porosities suitable for bone tissue
engineering are reviewed. Different polymer and ceramic compositions applied and their impact on biodegradability and bioactivity of
the scaffolds are discussed, including in vitro and in vivo assessments. The mechanical properties of todays available porous scaffolds are
analyzed in detail, revealing insufcient elastic stiffness and compressive strength compared to human bone. Further challenges in
scaffold fabrication for tissue engineering such as biomolecules incorporation, surface functionalization and 3D scaffold characterization
are discussed, giving possible solution strategies. Stem cell incorporation into scaffolds as a future trend is addressed shortly, highlighting
the immense potential for creating next-generation synthetic/living composite biomaterials that feature high adaptiveness to the
biological environment.
r 2006 Elsevier Ltd. All rights reserved.
Keywords: Scaffolds; Bioactivity; Bone tissue engineering; Composites; Porosity; Biodegradability
Contents
1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3414
2. Biodegradable polymer matrices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3414
2.1. Saturated aliphatic polyesters (PLA, PGA and PCL). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3415
2.2. Polypropylene fumarate (PPF). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3416
2.3. Polyhydroxyalkanoates (PHB, PHBV, P4HB, PHBHHx, PHO) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3416
2.4. Surface bioeroding polymers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3417
3. Bioactive ceramic phases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3417
3.1. Bioactive glasses and glass-ceramics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3418
3.1.1. Composition and bioactivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3418
3.1.2. Mechanical properties. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3419
3.2. Calcium phosphates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3419
3.2.1. Composition and bioactivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3419
3.2.2. Mechanical properties. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3419
3.3. Other bioactive ceramics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3419
ARTICLE IN PRESS
www.elsevier.com/locate/biomaterials
0142-9612/$ - see front matter r 2006 Elsevier Ltd. All rights reserved.
doi:10.1016/j.biomaterials.2006.01.039

Corresponding author. Department of Materials, Imperial College London, Prince Consort Road, London SW7 2BP, UK. Tel.: +44 207 594 6731;
fax: +44 207 594 6757.
E-mail address: a.boccaccini@imperial.ac.uk (A.R. Boccaccini).
4. Material processing strategies for composite scaffolds with interconnected pores . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3420
4.1. Thermally induced phase separation (TIPS) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3421
4.2. Solvent casting and particle leaching . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3423
4.3. Solid freeform fabrication techniques (SFFT) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3423
4.4. Microsphere sintering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3423
4.5. Coated scaffolds. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3424
5. Challenges and opportunities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3424
5.1. Mechanical integrity of porous scaffolds. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3424
5.2. Incorporation of biomolecules . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3424
5.3. Long-term characterization of porous composite scaffolds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3425
5.4. In vitro and in vivo characterization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3425
6. Summary of current status and future trends . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3425
6.1. Summary. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3425
6.2. Future trend: stem cells and ideal biomaterials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3426
Acknowledgment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3426
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3426
1. Introduction
Tissue engineering applies methods from materials
engineering and life sciences to create articial constructs
for regeneration of new tissue [1]. One common approach
is to isolate specic cells through a small biopsy from a
patient to grow them on a three-dimensional (3D) scaffold
under controlled culture conditions. Subsequently, the
construct is delivered to the desired site in the patients
body with the aim to direct new tissue formation into the
scaffold that can be degraded over time [13]. An
alternative approach is to implant scaffolds for tissue
ingrowth directly in vivo with the purpose to stimulate and
to direct tissue formation in situ [2,4,5]. The advantage of
this approach is the reduced number of operations needed,
resulting in a shorter recovery time for the patient.
Facing a complex biological and sensitive system as the
human body, the requirements of scaffold materials for
tissue engineering are manifold and extremely challenging.
First, biocompatibility of the substrate materials is
imperative; that is the material must not elicit an
unresolved inammatory response nor demonstrate im-
munogenicity or cytotoxicity. In addition, the mechanical
properties of the scaffold must be sufcient and not
collapse during handling and during the patients normal
activities. As with all materials in contact with the human
body, tissue scaffolds must be easily sterilizable to prevent
infection [6]. This applies notably for bulk degradable
scaffolds, where both the surface and the bulk material
must be sterile. A further requirement for a scaffold
particularly in bone engineering is a controllable inter-
connected porosity to direct the cells to grow into the
desired physical form and to support vascularization of the
ingrown tissue. A typical porosity of 90% as well as a pore
diameter of at least 100 mm is known to be compulsory for
cell penetration and a proper vascularization of the
ingrown tissue [710]. Other highly desirable features
concerning the scaffold processing are near-net-shape
fabrication and scalability for cost-effective industrial
production.
Today, materials used for scaffolds are natural or
synthetic polymers such as polysaccharides, poly(a-hydro-
xy ester), hydrogels or thermoplastic elastomers [2,4,11,12].
Other important categories of materials are bioactive
ceramics such as calcium phosphates and bioactive glasses
or glass-ceramics [8,13,14]. Currently, composites of
polymers and ceramics are being developed with the aim
to increase the mechanical scaffold stability and to improve
tissue interaction [1419]. In addition, efforts have also
been invested in developing scaffolds with a drug-delivery
capacity. These scaffolds can locally release growth factors
or antibiotics and enhance bone ingrowth to treat bone
defects and even support wound healing [14,2023].
Aforementioned requirements for scaffold materials are
numerous. To fulll as many requirements as possible,
composite systems combining advantages of polymers and
ceramics seem to be a promising choice, in particular for
bone tissue engineering, as demonstrated by the increasing
research efforts worldwide [2,1422,2429]. This paper
reviews tissue engineering relevant biodegradable polymers
and bioactive ceramics, including strategies for fabrication
of composite scaffolds with interconnected pores. Micro-
structure and mechanical properties will be discussed and
compared, evaluating open challenges in this eld of
biomedical materials research. In vitro and in vivo
characteristics of porous composite scaffolds, with focus
on bone regeneration, will be discussed as well as
summarizing the currently available literature and pointing
to research and development needs.
2. Biodegradable polymer matrices
There are two types of biodegradable polymers: The
natural-based materials are one category, including poly-
saccharides (starch, alginate, chitin/chitosan, hyaluronic
acid derivatives) or proteins (soy, collagen, brin gels, silk)
and, as reinforcement, a variety of biobers such as
lignocellulosic natural bers which are described in detailed
studies and reviews elsewhere [3034].
ARTICLE IN PRESS
K. Rezwan et al. / Biomaterials 27 (2006) 34133431 3414
This review will focus on the second category, synthetic
biodegradable polymers. Synthetic polymers can be pro-
duced under controlled conditions and therefore exhibit in
general predictable and reproducible mechanical and
physical properties such as tensile strength, elastic modulus
and degradation rate. A further advantage is the control of
material impurities. Possible risks such as toxicity, im-
munogenicity and favoring of infections are lower for pure
synthetic polymers with constituent monomeric units
having a well-known and simple structure. Table 1 gives
an overview of the discussed polymers and their physical
properties [31,3544].
2.1. Saturated aliphatic polyesters (PLA, PGA and PCL)
The most often utilized biodegradable synthetic poly-
mers for 3D scaffolds in tissue engineering are saturated
poly-a-hydroxy esters, including poly(lactic acid) (PLA)
and poly(glycolic acid) (PGA), as well as poly(lactic-co-
glycolide) (PLGA) copolymers [2,31,45,46]. PLA exists in
three forms: L-PLA (PLLA), D-PLA (PDLA), and racemic
mixture of D,L-PLA (PDLLA).
The chemical properties of these polymers allow hydro-
lytic degradation through de-esterication. Once degraded,
the monomeric components of each polymer are removed
by natural pathways. The body already contains highly
regulated mechanisms for completely removing monomeric
components of lactic and glycolic acids. PGA is converted
to metabolites or eliminated by other mechanisms, and
PLA can be cleared through the tricarboxylic acid cycle.
Due to these properties PLA and PGA have been used in
products and devices, such as degradable sutures which
have been approved by the US Food and Drug Adminis-
tration [2]. PLA and PGA can be processed easily and their
degradation rates, physical and mechanical properties are
adjustable over a wide range by using various molecular
weights and copolymers. However, these polymers undergo
a bulk erosion process such that they can cause scaffolds to
fail prematurely. In addition, abrupt release of these acidic
degradation products can cause a strong inammatory
response [47,48].
In general, PGA degrades faster than PLA, as found in
Table 1. Their degradation rates decrease in the following
order:
PGA4PDLLA4PLLA4PCL:
Biodegradable polyester degradation occurs by uptake
of water followed by the hydrolysis of ester bonds.
Different factors affect the degradation kinetics, such as:
chemical composition and congurational structure, pro-
cessing history, molar mass (Mw), polydispersity (Mw/
Mn), environmental conditions, stress and strain, crystal-
linity, device size, morphology (e.g. porosity) and chain
orientation, distribution of chemically reactive compounds
within the matrix, additives [49,50], presence of original
monomers and overall hydrophilicity. PLGA, for instance,
has a wide range of degradation rates, the degradation
kinetics being governed by both hydrophobic/hydrophilic
balance and crystallinity. Composition of chains (i.e.
contents in L-LA and D-LA and/or GA units) determines
the degradation rate of PLGA polymers. Blends containing
the greatest amount of PGA have been shown to degrade
faster [49]. Poly(e-caprolactone) (PCL) on the other hand,
can take several years to degrade in vivo [38,51].
Thick samples of these polymers can lead to hetero-
geneous degradation, faster inside than at the exterior.
ARTICLE IN PRESS
Table 1
Physical properties of synthetic, biocompatible, and biodegradable polymers used as scaffold materials
Polymer Melting point T
m
(1C) Glass transition point
T
g
(1C)
Biodegradation time
(months)
Compressive* or
tensile strength (MPa)
Modulus (GPa)
1. Bulk degradable polymers
PDLLA [3638] Amorphous 5560 1216 Pellet: 35150
*
Film or disk: 1.92.4
Film or disk: 2935
PLLA [36,38] 173178 6065 424 Pellet: 40120* Film or disk: 1.23.0
Film or disk: 2850
Fibre: 8702300 Fibre: 1016
PGA [3739] 225230 3540 612 Fibre: 340920 Fibre: 714
PLGA [31] Amorphous 4555 Adjustable: 112 41.455.2 1.42.8
PPF [31,40] Bulk 230
*
PCL [41] 58 72 424
PHA and blends [62] 120177 2 to 4 Bulk 2043
2. Surface erodative polymers
Poly(anhydrides)
[31,40,41]
150200 Surface 2527 0.141.4
3040
*
Poly(ortho-esters)
[31,42]
30100 Surface 416
*
2.54.4
Polyphosphazene [43] 66 to 50 242 Surface
K. Rezwan et al. / Biomaterials 27 (2006) 34133431 3415
Heterogeneous degradation may be ascribed to two
phenomena [46]:
(i) easier diffusion of soluble oligomers from the
surface into the external medium than from inside,
and
(ii) neutralization of carboxylic end groups located at
the surface by the external buffer solution (in vitro
or in vivo).
These phenomena contribute to reduce the acidity at
the surface whereas, in the bulk, degradation rate is
enhanced by autocatalysis due to carboxylic end groups.
Hydrolysis of amorphous polymers, e.g. PDLLA, is
faster due to the lack of crystalline regions. In general,
the amount of absorbed water depends on diffusion
coefcients of chain fragments within the polymer matrix,
temperature, buffering capacity, pH, ionic strength, addi-
tions in the matrix, in the medium and processing history.
Different aliphatic polyesters can therefore exhibit quite
distinct degradation kinetics in aqueous solutions.
PGA, for example, is a stronger acid and is more
hydrophilic than PLA, which is hydrophobic due to its
methyl groups.
The stereochemistry inuences the nal properties; better
alignment of neighbors leads to higher crystallinity. In
general, the initial degree of crystallinity of polyesters
affects the rate of hydrolytic degradation, as the crystal
segments are chemically more stable than amorphous
segments and reduce water permeation into the matrix.
Degradation takes longer with the stereoisomers of the
polymer, e.g. PLA composed of L-lactic repeating units
takes more than 5 years for total absorption, whereas only
about 1 year is needed for amorphous PLA (or PDLLA)
[51].
Of particular signicance for applications in tissue
engineering are debris and crystalline by-products, as well
as particularly acidic degradation products of PLA, PGA,
PCL and their copolymers that have been implicated in
adverse tissue reactions [18,38]. Several groups have
incorporated basic compounds to stabilize the pH of the
environment surrounding the polymer and to control its
degradation. Bioactive glasses and calcium phosphates
have been used [4951]. In fact, the possibility of counter-
acting the acidic degradation of biodegradable polymers is
another reason given for the use of composites [52] as
discussed further below.
PDLLA has been extensively investigated as a biomedi-
cal coating orthopedic material because of its excellent
features with respect to implant performance [31,53]. In
addition to its high mechanical stability [54], PDLLA also
shows excellent biocompatibility in vivo and a good
osteoconductive potential [55]. PDLLA of low molecular
weight can be combined with drugs like growth factors,
antibiotics, or thrombin inhibitor [56] to establish a locally
acting drug-delivery system. It is because of these desirable
features that much more attention has recently been paid to
PDLLA for applying it as a scaffold material for tissue
engineering.
PCL is also an important member of the aliphatic
polyester family. It has been used to effectively entrap
antibiotic drugs and thus a construct made with PCL can
be considered as a drug-delivery system, being used to
enhance bone ingrowth and regeneration in the treatment
of bone defects [57]. The degradation of PCL and its
copolymers involves similar mechanisms to that of PLA,
proceeding in two stages: random hydrolytic ester cleavage
and weight loss through the diffusion of oligometric species
from the bulk. It has been found that the degradation of
PCL system with a high molecular weight (

M
n
of 50,000) is
remarkably slow, requiring 3 years for complete removal
from the host body [58].
2.2. Polypropylene fumarate (PPF)
PPF is an unsaturated linear polyester. Like PLA and
PGA, the degradation products of PPF (i.e. propylene
glycol and fumaric acid) are biocompatible and readily
removed from the body. The double bond along the
backbone of the polymer permits cross-linking in situ,
which causes a moldable composite to harden within
1015 min. Mechanical properties and degradation time of
the composite may be controlled by varying the PPF
molecular weight. Therefore, preservation of the double
bonds and control of molecular weight during PPF
synthesis are critical issues [59]. PPF has been suggested
for use as a scaffold for guided tissue regeneration, often as
part of an injectable bone replacement composite [60]. It
also has been used as a substrate for osteoblast cultures
[61]. The development of composite materials combining
PPF and inorganic particles, e.g. hydroxyapatite (HA) or
bioactive glasses, has not been investigated to a large
extent, in comparison with the extensive research efforts
dedicated to PLGA- and PLA-based composites.
2.3. Polyhydroxyalkanoates (PHB, PHBV, P4HB,
PHBHHx, PHO)
Polyhydroxyalkanoates (PHA) are aliphatic polyesters
as well, but produced by microorganisms under unba-
lanced growth conditions [62,63]. They are generally
biodegradable (via hydrolysis) and thermoprocessable,
making them attractive as biomaterials for applications in
medical devices and tissue engineering. Over the past years,
PHA, particularly poly-3-hydroxybutyrate (PHB), copoly-
mers of 3-hydroxybutyrate and 3-hydroxyvalerate (PHBV),
poly-4-hydroxybutyrate (P4HB), copolymers of 3-hydro-
xybutyrate and 3-hydroxyhexanoate (PHBHHx) and poly-
3-hydroxyoctanoate (PHO) were demonstrated to be
suitable for tissue engineering and are reviewed in detail
in Ref. [35].
Dependent on the property requirement by different
applications, PHA polymers can be either blended, surface
modied or composed with other polymers; enzymes or
ARTICLE IN PRESS
K. Rezwan et al. / Biomaterials 27 (2006) 34133431 3416
inorganic materials to further adjust their mechanical
properties or biocompatibility. The blending among the
several PHA themselves can change dramatically the
material properties and biocompatibility [35,62].
PHB is of particular interest for bone tissue application
as it was demonstrated to produce a consistent favorable
bone tissue adaptation response with no evidence of an
undesirable chronic inammatory response after implanta-
tion periods of up to 12 months. Bone is formed close to
the material and subsequently becomes highly organized,
with up to 80% of the implant surface lying in direct
apposition to new bone. The materials showed no evidence
of extensive structural breakdown in vivo during the
implantation period of the study [64].
However, a drawback of some PHA polymers is their
limited availability and the time-consuming extraction
procedure from bacterial cultures that is required for
obtaining sufcient processing amounts as described in the
literature [35,65]. Therefore, the extraction process might
be a challenge to a cost-effective industrial upscale
production for large amounts of some PHA polymers.
2.4. Surface bioeroding polymers
There is a family of polymers that undergoes a
heterogeneous hydrolysis process which is predominantly
conned to the polymerwater interface. This property is
referred to as surface eroding as opposed to bulk
degrading behavior. Polymers known to show this
property are poly(anhydrides), poly(ortho-esters) and
polyphosphazene. These surface bioeroding polymers have
been intensively investigated as drug-delivery vehicles. The
surface-eroding characteristics offer three key advantages
over bulk degradation when used as scaffold materials: (1)
retention of mechanical integrity over the degradative
lifetime of the device, owing to the maintenance of mass to
volume ratio, (2) minimal toxic effects (i.e. local acidity),
owing to lower solubility and concentration of degradation
products, and (3) signicantly enhanced bone ingrowth
into the porous scaffolds, owing to the increment in pore
size as the erosion proceeds [66]. This group of polymers
can also be designed to be bulk degradable by introducing
a high surface to bulk ratio to the scaffold. Their properties
are summarized in Table 1.
3. Bioactive ceramic phases
A common characteristic of bioactive glasses and
ceramics is a time-dependent kinetic modication of the
surface that occurs upon implantation. The surface forms a
biologically active hydroxy carbonate apatite (HCA) layer
which provides the bonding interface with tissues. The
HCA phase that forms on bioactive implants is chemically
and structurally equivalent to the mineral phase in bone,
providing interfacial bonding [13,67]. The in vivo forma-
tion of an apatite layer on the surface of a bioactive
ceramic can be reproduced in a protein-free and acellular
simulated body uid (SBF), which is prepared to have an
ARTICLE IN PRESS
Table 2
Mechanical properties of dense and highly porous hydroxyapatite, 45S5 Bioglass
s
, A/W glass-ceramic, and human cortical bone
Ceramics Compressive strength
(MPa)
Tensile strength
(MPa)
Elastic modulus (GPa) Fracture toughness
(Mpa

m
p
)
References
Hydroxyapatite (HA) 4400 40 100 1.0 [70,71]
45S5 Bioglass
s
500 42 35 0.51 [71,72,76]
Glass-ceramicA/W 1080 215 118 2.0 [73]
Porous bioactive glass70S30C
(82%)
2.25 [105]
Porous Bioglass
s
-derived glass-
ceramic (490%)
0.2-0.4 [69]
Porous HA (8286%) 0.210.41 0.831.6 10
3
[106]
Cortical bone 130180 50151 1218 68 [31,74,75]
Cancellous bone 412 0.10.5 [107,108]
Fig. 1. SEM micrographs illustrating the typical cauliower morphol-
ogy of hydroxyapatite formed on the surface of a 45S5 Bioglass
s
-based
foam after immersion in simulated body uid (SBF) for 28 days. The foam
was sintered at 1000 1C for 1 h. The magnication of the framed area
shown in the inlet picture reveals rod-shaped crystals of hydroxyapatite
(adapted from Ref. [69]).
K. Rezwan et al. / Biomaterials 27 (2006) 34133431 3417
ion concentration nearly equal to that of human blood
plasma. The composition of SBF can be found in Ref. [68].
As an example, the typical cauliower morphologies of
HCA layers formed on a scaffold made of 45S5Bioglass
s
immersed in SBF are illustrated by SEM micrographs in
Fig. 1 (adapted from Ref. [69]). Typical mechanical
properties of the bioactive ceramic phases discussed in
the following paragraphs can be found in Table 2,
compiled from available literature data [31,7076].
3.1. Bioactive glasses and glass-ceramics
In 1969, Hench et al. discovered that certain glass
compositions had excellent biocompatibility as well as the
ability of bone bonding [77]. Through interfacial and cell-
mediated reactions, bioactive glass develops a calcium-
decient, carbonated phosphate surface layer that allows it
to chemically bond to host bone. This bone-bonding
behavior is referred to as bioactivity and has been
associated with the formation of a carbonated hydroxya-
patite (HCA) layer on the glass surface when implanted
or in contact with biological uids [13,71,7680] (see also
Fig. 1).
The stages that are involved in forming the bone bond of
bioactive glasses and bioactive glass-ceramics were sum-
marized by Hench [13,67]. Although some details remain
unknown, it is clearly recognized that for a bond with bone
tissue to occur, a layer of biologically active HCA must
form. This conclusion is based on the nding that HCA is
the only common characteristic of all the known bioactive
implant materials [81]. Bioactivity, however, is not an
exclusive property of bioactive glasses. HA and related
calcium phosphates also show an excellent ability to bond
to bone, as discussed further below. The capability of a
material to form a biological interface with surrounding
tissue is critical in elimination of scaffold loosening. Of
great importance and impact for applications in tissue
engineering is that bioactive glasses have also been found
to support enzyme activity [8285]; vascularization [86,87];
foster osteoblast adhesion, growth, differentiation; and
induce the differentiation of mesenchymal cells into
osteoblasts [26,8890].
A signicant nding for the development of bone
engineering is that the dissolution products from bioactive
glasses, in particular the 45S5 Bioglass
s
composition,
upregulate the gene expression that control osteogenesis
and the production of growth factors [91]. Silicon has been
found to play a key role in the bone mineralization and
gene activation, which has led to an increased interest in
the substitution of silicon for calcium into synthetic HA.
Investigations in vivo have shown that bone ingrowth into
silicon-substituted HA granules was remarkably greater
than that into pure HA [92].
The above-mentioned advantages are the reasons why
45S5 Bioglass
s
is successfully used in clinical treatments of
periodontal disease (Perioglas
TM
) and as a bone ller
material (Novabone
TM
) [13]. Bioglass
s
implants have also
been used to replace damaged middle ear bones, restoring
hearing to patients [79]. Bioactive glasses have gained new
attention recently as promising scaffold materials, either as
ller or coatings of polymer structures, or as porous
materials themselves, which involves melt-derived and
solgel-derived glasses [52,69,9399].
3.1.1. Composition and bioactivity
The basic constituents of the most bioactive glasses are
SiO
2
, Na
2
O, CaO, and P
2
O
5
. 45S5 Bioglass
s
contains 45%
SiO
2
, 24.5% Na
2
O, 24.4% CaO and 6% P
2
O
5
, in weight
percent. An overview of different bioactive glass composi-
tions and their corresponding bioactivities are given in
Ref. [67]. Hench and coworkers have systematically studied
a series of glasses in the four-component systems with
a constant 6 wt% P
2
O
5
content, as summarized in
Refs. [67,81] and they divided the compositions into three
regions according to their bioactivity. Bioactive glasses
(e.g. 45S5 Bioglass
s
) with compositions in the system
SiO
2
Na
2
OCaOP
2
O
5
, having o55% SiO
2
, exhibit a
high bioactivity index (in region A), and bond to both
soft and hard connective tissues. The bioactive glasses
(i.e. glasses in region A) are osteoproductive (bone
grows on material surfaces due to enhanced osteoblast
activity) and osteoconductive. Glasses of compositions
in region B exhibit only osteoconductivity. Compositions
in region C are resorbed within 1030 days in tissue
[13].
It has been found that reactions on bioactive glass
surfaces can release critical concentrations of soluble Si,
Ca, P and Na ions, depending on the processing route and
particle size. The released ions induce intracellular and
extracellular responses [91,100]. For example, a synchro-
nized sequence of genes is activated in osteoblasts that
undergo cell division and synthesize an extracellular
matrix, which mineralizes to become bone. In addition,
bioactive glass compositions doped with AgO
2
have been
shown to elicit anti-bactericidal properties while maintain-
ing their bioactive function [101]. In recent investigations,
45S5 Bioglass
s
has also been shown to increase secretion
of vascular endothelial growth factor (VEGF) in vitro and
to enhance vascularization in vivo, suggesting scaffolds
containing controlled concentrations of Bioglass
s
might
stimulate neo-vascularization which is benecial to large
tissue engineered constructs [86].
One key reason that makes bioactive glasses a relevant
scaffold material is the possibility of controlling a range of
chemical properties and thus the rate of bioresorption. The
structure and chemistry of glasses, in particular solgel-
derived glasses [78,79], can be tailored at a molecular level
by varying either composition, or thermal or environ-
mental processing history. It is possible to design glasses
with degradation properties specic to a particular
application of bone tissue engineering.
However, it was reported that crystallization of bioactive
glasses decreased the level of bioactivity [102] and even
turned a bioactive glass into an inert material [103]. This is
ARTICLE IN PRESS
K. Rezwan et al. / Biomaterials 27 (2006) 34133431 3418
a disadvantage that limits the application of bioactive
glasses as scaffold materials, as full crystallization happens
prior to densication by viscous ow sintering [104].
Extensive sintering however is necessary to densify the
struts of a scaffold to achieve the required mechanical
stability.
3.1.2. Mechanical properties
A drawback of bioactive glasses is their low fracture
toughness and mechanical strength, especially in a porous
form. Hence, bioactive glasses alone have limited applica-
tion in load-bearing situations. Table 2 gives compressive
strength, elastic modulus and some fracture toughness
values for dense and porous bioactive ceramics, compiled
from Refs. [31,6975,105108]. It can be seen that
particularly porous scaffolds needed for tissue engineering
exhibit very low mechanical properties compared to
cortical and cancellous bone.
Bioactive glass-ceramic materials can exhibit better
mechanical properties than amorphous glass and calcium
phosphate ceramics (Table 2). The high bending strength of
apatite-wollastonite (A/W) glass-ceramic is due to the
precipitation of the wollastonite as well as apatite phases
and attributed to the high fracture toughness of the
precipitation microstructure [109].
3.2. Calcium phosphates
3.2.1. Composition and bioactivity
Around 60 wt% of bone is made of HA Ca
10
(PO
4
)
6
(OH)
2
and therefore it is evident why HA and
related calcium phosphates (e. g. a-TCP, b-TCP) have been
intensively investigated as the major component of scaffold
materials for bone tissue engineering [76,110112]. As
expected, calcium phosphates have an excellent biocompat-
ibility due to their close chemical and crystal resemblance
to bone mineral [113]. Although they have not shown
osteoinductive ability, calcium phosphates certainly pos-
sess osteoconductive properties and may bind directly to
bone under certain conditions [114116]. Numerous in vivo
and in vitro assessments have reported that calcium
phosphates, no matter of which form (bulk, coating,
powder, or porous) and of which phase (crystalline or
amorphous), always support the attachment, differentia-
tion, and proliferation of relevant cells (such as osteoblasts
and mesenchymal cells), with HA being possibly the most
efcient among them [117]. While the excellent biological
performance of HA and related crystalline calcium
phosphates has been well documented, their relatively slow
biodegradation and in particular low mechanical strength
limit their application in engineering of new bone tissue,
especially at load-bearing sites.
Crystalline calcium phosphates have long degradation
times in vivo, typically in the order of months or even
years. The dissolution rate of synthetic HA depends on the
type and concentration of the buffered or unbuffered
solutions, pH of the solution, degree of the saturation of
the solution, solid/solution ratio, and the composition and
crystallinity of the HA phase. In the case of crystalline HA,
the degree of micro- and macro-porosities, defect structure
and amount and type of other phases present have also a
signicant inuence [116]. Crystalline HA exhibits the
slowest degradation rate, compared with other calcium
phosphates. The dissolution rate decreases in the following
order [118]:
Amorphous HAba TCPbb TCPb crystalline HA:
3.2.2. Mechanical properties
In the body, the mechanical properties of natural bone
change with their biological location because the crystal-
linity, porosity, and composition of bone adjust to the
biological and biomechanical environment. The properties
of synthetic calcium phosphates vary signicantly with
their crystallinity, grain size, porosity, and composition
(e.g. calcium deciency) as well. In general, the mechanical
properties of synthetic calcium phosphates decrease sig-
nicantly with increasing amorphous phase, microporosity
and grain size. High crystallinity, low porosity and small
grain size tend to give higher stiffness, compressive and
tensile strengths, and greater fracture toughness. It has
been reported that the exural strength and fracture
toughness of dense HA are much lower in a dry condition
than in a wet condition [119].
If we compare the properties of HA and related calcium
phosphates with those of cortical bone (Table 2), we nd
that bone has a reasonably good compressive strength
though it is lower than that of HA. But bone has a
signicantly higher fracture toughness than HA. The
mechanical properties are even lower for porous HA
structures. The high tensile strength and fracture toughness
of bone are attributed to the tough and exible collagen
bers reinforced by HA crystals. Hence, calcium
phosphates alone cannot be used for load-bearing
scaffolds despite their good biocompatibility and osteo-
conductivity.
3.3. Other bioactive ceramics
Representative bioactive ceramics are, as mentioned,
Bioglass
s
, HA and glass-ceramics containing HA or its
components, such as CaO and P
2
O
5
. However, formation
of HA does not seem to be limited to those ceramics only.
As Kokubo et al. [68] reported, other materials can show
degrees of bioactivity after a simple chemical heat
treatment as well. Chemical treatment of metals and
ceramics, e.g. by NaOH and heat treatments of titanium
metal, titanium alloys, and tantalum metal, and by H
3
PO
4
treatment of tetragonal zirconia, resulted in functional
graded surfaces that induced HA formation on the surface
in SBF studies. In vivo studies of NaOH and heat-treated
titanium metals implanted in a rabbit femur showed
detaching fracture loads of up to 4 times higher than the
untreated implants [68].
ARTICLE IN PRESS
K. Rezwan et al. / Biomaterials 27 (2006) 34133431 3419
It has been suggested that the TiOH, ZrOH, NbOH,
and TaOH surface groups in anatase or the tetragonal/
monoclinic structures may provide effective epitaxial
nucleation sites for apatite crystals and that a negative
surface charge increases Ca
2+
adsorption from the
solution. In addition, the chemical treatment induces the
incorporation of ions (e.g. Na
+
) on the metal oxide surface
and transforms the crystal phase to an amorphous phase
resulting in an increased ion solubility [120124]. However,
the suggested models have not been investigated in detail
yet and more studies will be required to elucidate the
mechanisms.
4. Material processing strategies for composite scaffolds
with interconnected pores
Development of composite scaffold materials is attrac-
tive as advantageous properties of two or more types of
materials can be combined to suit better the mechanical
and physiological demands of the host tissue. By taking
advantage of the formability of polymers and including
controlled-volume fractions of a bioactive ceramic phase,
mechanical reinforcement of the fabricated scaffold can be
achieved [52,125]. At the same time, the poor bioactivity of
most polymers can be counteracted.
Probably the most important driving force behind the
development of polymer/bioactive glass composite scaf-
folds for bone tissue engineering is the need for conferring
bioactive behavior to the polymer matrix, which is
achieved by the bioactive inclusions or coatings. The
degree of bioactivity is adjustable by the volume fraction,
size, shape and arrangement of inclusions [24,52,126137].
It has been shown that increased volume fraction and
higher surface area to volume ratio of inclusions favor
higher bioactivity, hence in some applications the incor-
poration of bers instead of particles is favored [27,138].
Addition of bioactive phases to bioresorbable polymers
can also alter the polymer degradation behavior, by
allowing rapid exchange of protons in water for alkali in
the glass or ceramic. This mechanism is suggested to
provide a pH buffering effect at the polymer surface,
modifying the acidic polymer degradation [24,52,139].
Inclusion of bioactive glasses has been shown to modify
surface and bulk properties of composite scaffolds by
increasing the hydrophilicity and water absorption of the
hydrophobic polymer matrix, thus altering the scaffold
degradation kinetics. In particular, the inclusion of 45S5
Bioglass
s
particles was found to increase water absorption
compared to pure polymer foams of PDLLA [133] and
PLGA [24,52]. In related research, it has been reported that
polymer composites lled with HA particles hydrolyzed
homogeneously due to water penetrating the interfacial
regions [140]. Ideally, the degradation and resorption
kinetics of composite scaffolds are designed to allow cells
to proliferate and secrete their own extracellular matrix
while the scaffolds gradually vanish, leaving space for
new cell and tissue growth. The physical support provided
by the 3D scaffold should be maintained until the
engineered tissue has sufcient mechanical integrity to
support itself.
There are numerous foaming techniques including solgel
routes to obtain highly porous structures [141,142]. How-
ever, only relevant fabrication techniques leading to 3D
composite scaffolds with highly interconnected pores are
discussed in the following paragraphs and compared in
Table 3, compiled with data available from the literature
[133,135,143148]. A selection of dense and porous scaffold
composites including their physical properties is given in
ARTICLE IN PRESS
Table 3
Fabrication routes for 3D composite scaffolds with high pore interconnectivity and their advantages and disadvantages
Fabrication route Advantages Disadvantages
Thermally induced phase separation
(TIPS) [133,143]
High porosities (95%) Long time to sublime solvent (48 hours)
Highly interconnected pore structures Shrinkage issues
Anisotropic and tubular pores possible Small scale production
Control of structure and pore size by varying
preparation conditions
Use of organic solvents
Solvent casting/particle leaching [144,145] Controlled porosity Structures generally isotropic
Controlled interconnectivity (if particles are
sintered)
Use of organic solvents
Solid free-form [146,147] Porous structure can be tailored to host tissue Resolution needs to be improved to the micro-scale
Protein and cell encapsulation possible Some methods use organic solvents
Good interface with medical imaging
Microsphere sintering [148] Graded porosity structures possible Interconnectivity is an issue
Controlled porosity Use of organic solvents
Can be fabricated into complex shapes
Scaffold coating [135] Quick and easy Clogging of pores, sometimes organic solvents used,
coating adhesion to substrate can be too weak
K. Rezwan et al. / Biomaterials 27 (2006) 34133431 3420
Table 4, which represent typical systems reported in
the literature [15,17,25,28,29,31,74,75,93,107,108,127,132,
134,136,148159].
For comparison, the mechanical properties of human
cortical and cancellous bone are listed as well. Representa-
tive morphologies of the fabricated scaffolds are illustrated
in Fig. 2, taken from Refs. [147,148,160162].
4.1. Thermally induced phase separation (TIPS)
3D resorbable polymer scaffolds with very high poros-
ities (97%) can be produced using the TIPS technique to
give controlled macro- and microstructures suitable as
scaffolds for tissues such as nerve, muscle, tendon,
ligament, intestine, bone, and teeth [52,143,163]. The
obtained scaffolds are highly porous with anisotropic
tubular morphology and extensive pore interconnectivity.
Microporosity of TIPS produced foams, their pore
morphology, mechanical properties, bioactivity and degra-
dation rates can be controlled by varying the polymer
concentration in solution, volume fraction of the secondary
phase, quenching temperature and the polymer and solvent
used as discussed in a previous review paper [52].
Briey, the polymer is dissolved in dimethylcarbonate
and stirred overnight to obtain a homogeneous polymer
solution. A given amount of glass or ceramic powder can
be added into the polymer solution. The mixture is
transferred into a ask and sonicated. Hereafter, the
ask is quenched in liquid nitrogen and maintained
at 196 1C for 2 h. The frozen mixture is then transferred
into a cooling bath at 10 1C and connected to a vacuum
pump. The solvent is sublimated at 10 1C for 48 h and
then at 0 1C for 48 h, followed by drying at room
temperature in a vacuum oven until reaching a constant
weight [24].
Maquet et al. [24,133] developed highly porous PDLLA/
Bioglass
s
composite scaffolds prepared by TIPS with
bimodal and anisotropic pore structures composed of
tubular macropores of 100 mm, interconnected with
micropores of 1050 mm in diameter, as shown in Fig. 2a.
The pore volume was shown to decrease from 9.5 to
5.7 cm
3
/g after including 40 wt% Bioglass
s
, with little
change observed in the overall pore morphology [164]. In
vitro studies in phosphate-buffered saline at 37 1C showed
that addition of Bioglass
s
increased water absorption and
weight loss in comparison to pure polymer foams [24,133].
The molecular weight was found to decrease less within the
composite foams, possibly due to the dissolution of
alkaline ions from the Bioglass
s
providing a pH buffering
effect, as discussed above. Both the PDLLA/Bioglass
s
composites and the neat PDLLA foams retained their
structural integrity until the end of the experiment (16
weeks), which means degradation was still in the early
stages.
ARTICLE IN PRESS
Table 4
Selection of scaffold composites designed for bone tissue engineering and their properties
Scaffold composite Percentage of
ceramic (%)
Porosity (%) Pore size (mm) Compressive (C),
tensile (T), exural
(F) strength (MPa)
Modulus (MPa) Reference
Ceramic Polymer
1. Dense composites
HA bre PDLLA 210.5 (vol.) 45 (F) 1.752.47 10
3
[149]
PLLA 1070 (wt.) 5060 (F) 6.412.8 10
3
[150]
HA PLGA 4085 (vol.) 22 (F) 1.1 10
3
[29,151]
b-TCP PLLA-co-PEH 75 (wt.) 51 (F) 5.18 10
3
[152,153]
PPF 25 (wt.) 7.57.7 (C) 191134
A/W PE 1050 (vol.) 1828 (B) 0.95.7 10
3
[154]
Cortical bone 50150(T) 1218 10
3
[31,74,75]
130180 (C)
2. Porous composites
Amorphous CaP PLGA 2875 (wt.) 75 4100 65 [28,155]
HA PLLA 50 (wt.) 8596 100 300 0.39 (C) 1014 [127]
PLGA 6075 (wt.) 8191 8001800 0.070.22 (C) 27.5 [156]
PLGA 3040 110150 3371459 [157]
Bioglass
s
PLGA 75 (wt.) 43 89 0.42 (C) 51 [93,148,158]
PLLA 2050 (wt.) 7780 100 (macro) 1.53.9 (T) 137260 [17]
10 (micro)
PLGA 0.11 (wt.) 50300 [15,148]
PDLLA 529 (wt.) 94 100 (macro) 0.070.08 0.651.2 [132,134,136]
1050 (micro)
Phosphate glass A/W PLA-PDLLA 40 (wt.) 9397
PDLLA 2040 (wt.) 85.595.2 98154 0.0170.020 (C) 0.0750.12 [159,25]
Cancellous Bone 412 (C) 100500 [107,108]
K. Rezwan et al. / Biomaterials 27 (2006) 34133431 3421
It is well documented [52,163,165] that due to auto-
catalysis non-porous biodegradable polylactides undergo
degradation more rapidly than the porous counterparts.
The reason for this effect lies in the fact that porous
materials are able to facilitate dissolving and spreading of
degradation products throughout the aqueous medium,
thereby preventing autocatalysis. PDLLA/Bioglass
s
com-
posites exhibit high bioactivity, assessed by the formation
of HA on the composite surfaces upon immersion in SBF
[133,134]. It has also been shown that the foams support
the migration, adhesion, spreading and viability of MG-63
cells (osteosarcoma cell line) [134].
The potential of these scaffolds in bone and soft-tissue
engineering has been demonstrated in vitro with optimized
concentrations of 45S5 Bioglass
s
added to PDLLA or
PLGA matrices [134,136]. Highly porous tubular scaffolds
with oriented porosity have also been fabricated by
exploiting the TIPS process [15,166]. These are candidate
materials for soft-tissue engineering with potential applica-
tion in the regeneration of tissues requiring tubular shapes
such as the intestine, trachea and blood vessels. TIPS
fabricated PDLLA foams with and without Bioglass
s
additions have been shown to exhibit mechanical aniso-
tropy concomitant with the TIPS-induced pore architecture
[132].
Polymer matrix composite lms containing nanosized
titania and other inorganic particulate inclusions have
demonstrated enhanced cell adhesion and a tendency
to increased Ca-containing mineral deposition [167].
Recently, 3D PDLLA foams containing both TiO
2
ARTICLE IN PRESS
Fig. 2. Typical morphologies of porous polymer foams produced by different techniques and structure of cancellous bone. (a) Thermal induced phase
separation (TIPS, adapted from Ref. [160]), (b) solvent casting and particle leaching [161], (c) solid freeform fabrication technique [147], (d) microsphere-
sintering [148], (e) cancellous bone [162]. (Micrographs (b), (c) and (e) reprinted with permission of Elsevier Ltd. Micrograph (d) reprinted with permission
of John Wiley & Sons, Inc.)
K. Rezwan et al. / Biomaterials 27 (2006) 34133431 3422
nanoparticles and Bioglass
s
additions have been synthe-
sized by TIPS. These foams demonstrate enhanced
bioactivity and surface nanotopography [131].
4.2. Solvent casting and particle leaching
Solvent casting of biocomposite scaffolds involves the
dissolution of the polymer in an organic solvent, mixing
with ceramic granules, and casting the solution into a
predened 3D mould. The solvent is subsequently allowed
to evaporate. The main advantage of this processing
technique is the ease of fabrication without the need of
specialized equipment. The primary disadvantages of
solvent casting are: (1) the limitation in the shapes
(typically at sheets and tubes are the only shapes that
can be formed); (2) the possible retention of toxic solvent
within the polymer; and (3) the denaturation of the
proteins and other molecules incorporated into the
polymer by the use of solvents. The use of organic solvents
to cast the polymer may decrease the activity of bioinduc-
tive molecules (e.g. protein). The detailed processing steps
have been described in the literature [93].
Polymer-ceramic constructs can also be fabricated by the
solvent aggregation method. The polymer microspheres are
rst formed from traditional water oil/water emulsions.
Solvent-aggregated polymer-ceramic scaffolds can then be
constructed by mixing solvent, salt or sugar particles, ceramic
granules, and pre-hardened microspheres [131]. A 3D structure
of controlled porosity is formed based on this method
combined with particle leaching and microsphere packing.
Fig. 2b illustrates a typical pore morphology obtained by this
technique. The method shares similar advantages and
disadvantages with the solvent casting technique [93].
There has been little work done on producing bioactive
polymer-ceramic scaffolds using particle leaching. Cer-
tainly, a drawback of this technique is achieving pore
interconnectivity at low porogen (salt/sucrose) loadings, as
many of the porogen particles may remain trapped.
Nevertheless, composites based on calcium phosphate
inclusions with variable and graded porosity have been
produced using this route [168].
4.3. Solid freeform fabrication techniques (SFFT)
SFFT, such as fused deposition modeling, have been
employed to fabricate highly reproducible scaffolds with
fully interconnected porous networks [147,169] as shown in
Fig. 2c. Using digital data produced by an imaging source
such as computer tomography or magnetic resonance
imaging enables accurate design of the scaffold structure
[169]. Solid freeform (SFF) manufacturing coupled with
conventional foam scaffold fabrication procedures (phase
separation, emulsion-solvent diffusion or porogen leach-
ing) may be used to develop scaffolds with controlled
micro- and macroporous structures. Such biomimetic
internal architectures may prove valuable for multi-tissue
and structural tissue interface engineering.
To the authors knowledge, there is no literature
available on degradable polymer/bioactive glass compo-
sites made by SFFT. This technique has been only applied
for composites containing calcium phosphates as the
bioactive phase [147,170]. For example, Xiong et al. [170]
fabricated composites of PLLA/TCP with porosities of up
to 90% and mechanical properties close to human
cancellous bone by using low-temperature deposition based
on a layer-by-layer manufacturing method of SFF fabrica-
tion (computer-driven by 3D digital models). PLLA was
dissolved in dioxane and TCP powder mixed to prepare a
slurry, which was formed into frozen scaffolds, and
subsequently freeze-dried. Alternate parallel layers formed
macropores (400 mm diameter) and sublimation of the
solvent during freeze-drying formed micropores (5 mm
diameter). Taboas et al. [147] produced PLA scaffolds
with computationally designed pores (500800 mm wide
channels) and solvent-derived local pores (50100 mm wide
voids or 510 mm length plates). Indirect fabrication using
casting in SFF moulds provided enhanced control over
scaffold shape, porosity and pore architecture, including
size, geometry, orientation, branching and interconnectivity.
A shortcoming of this route is increased scaffold fabrication
time compared with direct methods, as a temporary mould
must be made rst.
4.4. Microsphere sintering
In this process, microspheres of a ceramic and polymer
composite are synthesized rst, using emulsion/solvent
evaporation technique. Sintering the composite micro-
spheres yields a 3D, porous scaffold [28,155]. 3D compo-
sites of degradable polymers and bioactive glass have been
produced by sintering composite microspheres by Lu et al.
[148]. Starting materials were PLAGA-Bioglass
s
compo-
site microspheres obtained through a wateroilwater
emulsion technique. Sintering of the microspheres into
cylindrical shapes resulted in a well-integrated intercon-
nected porous structure, with the microspheres joined at
the contact necks. Average porosity was 40% with pore
diameters of 90 mm, and mechanical properties close to
cancellous bone. The composites were shown to be
bioactive as a calcium phosphate layer formed on the
surface of the composite on immersion in SBF for 7 days.
Moreover, Bioglass
s
reinforcement gave a two-fold
increase in compressive strength. The scaffolds were shown
to support the adhesion, growth and mineralization of
human osteoblast-like cells in vitro. Over a 3-week period,
cultures with PLAGA/Bioglass
s
maintained pH variations
within physiological ranges. More recently, Yao et al. [19]
synthesized PLGA/bioactive glass microspheres by emulsi-
cation and heated them in moulds to fabricate porous 3D
scaffolds. They demonstrated the bioactivity of the
composites and their ability to promote osteogenesis of
marrow stromal cells. A typical microsphere-sintered 3D
structure fabricated by Lu et al. [148] is given by Fig. 2d.
ARTICLE IN PRESS
K. Rezwan et al. / Biomaterials 27 (2006) 34133431 3423
4.5. Coated scaffolds
Bioceramic-coated porous scaffolds have been produced
either as foams [135], brous bodies [171] or meshes [86,158]
by slurry dipping or electrophoretic deposition (EPD).
Roether et al. [135] were the rst to develop composites of
macroporous PDLLA foams coated with Bioglass
s
parti-
cles (grade 45S5 with particle size o5 mm) by slurry dipping.
A stable and homogeneous coating on the foam surface and
inltration of Bioglass
s
particles throughout the porous
network were achieved. A stable slurry of 42 wt%
Bioglass
s
in deionized water gave relatively dense and
uniform adherent coatings. EPD was investigated as an
alternative route whereby charged Bioglass
s
particles in
aqueous suspension inltrated the foam with its tubular
macropores oriented perpendicularly to the larger dimen-
sion of the electrodes [135]. The slurry dipping technique
was found to be more suitable than the EPD route; the
latter caused frequently sealing of the interconnected pores
by Bioglass
s
particles. EPD, however, is an attractive
method to incorporate nanoparticles into porous structures
with potential use as tissue engineering scaffolds [172].
Composites tested in vitro in acellular SBF exhibited
increasing formation of HA (layers of 10 mm were formed
after 28 days). In addition, changes in pore morphology as
a result of polymer degradation with increasing immersion
time were observed. The investigation of the in vitro
behavior of osteoblast-like cells demonstrated that cells
were able to migrate through the porous network and
colonize the inner sections of the foams. Also, after 24 h a
higher cell density was observed in the Bioglass
s
coated
foams compared to the pure PDLLA foams [90].
A related approach, but using PHA woven meshes as
substrate, was followed by Olsen et al. [173]. In this case, a
45S5 Bioglass
s
aqueous slurry was used to deposit
micrometer-sized Bioglass
s
particles on the surfaces of
PHA bers. The process was optimized to coat individual
bers maintaining the pore structure of the woven mesh.
There are however no published results on the in vivo
behavior of these Bioglass
s
-PHA composites.
Another promising method to coat polymer surfaces with
bone like calcium phosphates is using a biomimetic approach
as demonstrated in the literature [174176]. Moreover,
sodium silicate gel [174] has been used to nucleate CaP
coatings on polymer surfaces that were immersed in SBF. A
review on biomimetic formation of calcium phosphate
coatings has been published recently [177].
5. Challenges and opportunities
5.1. Mechanical integrity of porous scaffolds
Comparison of the mechanical properties of todays
available porous scaffolds with relevant properties of bone
reveals the insufcient mechanical integrity of the man-
made scaffolds. In Fig. 3 the elastic modulus and the
compressive strength of dense bioactive ceramic, biode-
gradable polymers, cancellous and cortical bone are
compared with porous monophasic scaffolds and compo-
sites thereof. Mechanical data for porous bioactive
ceramics and for polymer foams were taken from Refs.
[105,106,132]. It can be seen that some dense polymers
match cancellous bone properties and approach cortical
bone properties. Moreover, the bioactive ceramics region is
close to the properties of cortical bone as well. Porous
scaffolds however are at least one order of magnitude
weaker than cancellous bone and orders of magnitude
weaker than cortical bone.
Interestingly, the stiffness achieved by up-to-date fabri-
cated porous bioactive ceramics is less than the stiffness of
most porous biodegradable polymer scaffolds. By compar-
ing the mechanical properties of the porous composites to
those of porous polymer scaffolds, a slight increase of
mechanical properties is revealed. But the increase in
stiffness and strength is certainly below expectations; most
probably this can be attributed to the lack of interfacial
bonding strength between the ceramic phase and the
polymer matrix, which has been neglected in most studies.
The ceramic phases are in general very hydrophilic whereas
the polymers are hydrophobic. The increase of bonding at
the interface is a challenge and might be achieved by using
surfactants chemisorbed on the particle surface prior to
composite processing. Using surface functionalized parti-
cles in the nanosize rangefeaturing a higher specic
surface and thus a higher interface areamight even
increase the interfacial bonding strength, and thus the
overall mechanical properties of the composite scaffold
could be effectively enhanced. However, the increase of
interfacial bonding and introduction of surfactants are
likely to have an impact on degradation kinetics and
cytotoxicity of the composite. These effects are largely
unknown and remain to be investigated.
5.2. Incorporation of biomolecules
There is a signicant scope in the application of surface
modications, through the use of protein adsorption or
ARTICLE IN PRESS
Fig. 3. Elastic modulus vs. compressive strength of biodegradable
polymers, bioactive ceramics and composites reviewed in this paper.
Porosities of the porous scaffolds are 475% and mostly interconnected.
K. Rezwan et al. / Biomaterials 27 (2006) 34133431 3424
plasma treatment, to provide more cues to cell attachment
and response [3,178]. The possibility of incorporating
growth factors into composites formed by biodegradable
polymers and bioactive glasses or HA inclusions has
started to be explored [179182].
Integrins, laminin and RGD proteins were shown to be
essential for cell attachment to materials surfaces
[183186]. The immobilization of these proteins should
not only promote cell adhesion and proliferation but also
increase wettability of hydrophobic polymers such as
PDLLA. To control protein adhesion and release kinetics,
different protein immobilization routes can be used as
demonstrated for polymer surfaces [187] and ceramic
surfaces [188,189]. Certain growth factors were shown in
in vivo studies to be osteoinductive. Possible growth
factors include bone morphogenetic proteins, transforming
growth factor beta, VEGF, and insulin-like growth factor
as reviewed in Refs. [190192]. Immobilizing these growth
factors on the scaffold surface might signicantly shorten
the bone healing process and reduce patient recovery time.
The incorporation of biomolecules does not allow
extreme temperature ranges (470 1C) or extremely aggres-
sive chemical conditions during processing, being a
challenge to the scaffold fabrication process. Soft
material routes like solgel processing might be a strategy
to incorporate biomolecules during scaffold fabrication. To
the authors knowledge, however, solgel-derived bioactive
organic/inorganic hybrids have not yet been formed into
highly interconnected porous structures, which is essential
for application of these composites as scaffolds. Another
related challenge is the elucidation of the local impact of
growth factors on the cell and tissue systems, including
long-term effects.
5.3. Long-term characterization of porous composite
scaffolds
There is a lack of current understanding in the literature
regarding the long-term in vitro and in vivo characteriza-
tion of the porous 3D scaffold composites discussed here,
specically regarding the long-term effect of the incorpora-
tion of inorganic bioactive phases on the degradation and
ion release kinetics of these highly porous systems. In this
regard, the development of appropriate characterization
techniques coupled with predictive analytical models is
mandatory in order to be able to comprehensively assess
the degradation of these systems with respect to pore
structure, scaffolds geometry, uid ow and the inuence
of the bioactive additions. Here, the use of X-ray
microtomography as a reliable tool for 3D pore structure
quantication is likely to gain increased impetus allowing
resolutions down to 1 mm [193]. Combining image analysis
and impedance spectroscopy is another possible approach
to characterize pore interconnectivities of scaffolds as
shown recently, being less straightforward however [164].
5.4. In vitro and in vivo characterization
While a good number of in vitro and in vivo studies exist
for biodegradable polymers and bioactive ceramics alone,
in vitro studies for polymer/ceramic composites have just
started [15,19,86,134,136,137,148,194,195]. Table 5 gives
examples of the types of composite scaffolds investigated in
vitro and the applied cell cultures. Very few composite
systems have been investigated in vivo up to date. More
research needs also to be directed at assessing the
suitability of the reviewed bioactive composite scaffolds
in soft-tissue engineering strategies, including further
investigations of the effect of dissolution products from
the bioactive phase on vascularization and in vivo new
tissue growth.
6. Summary of current status and future trends
6.1. Summary
The synthetic and biodegradable, polymer/inorganic
bioactive phase composites reviewed in this article are
particularly attractive as tissue engineering scaffolds due to
their shapability, bioactive behavior and adjustable biode-
gradation kinetics. Conventional materials processing
methods have been adapted and extended for incorpora-
tion of inorganic bioactive phases into porous and
interconnected 3D polymer networks.
From the materials science perspective, the present
challenge in tissue engineering is to design and fabricate
reproducible bioactive and bioresorbable 3D scaffolds of
tailored porosity and pore structure, which are able to
maintain their structure and integrity for predictable times,
even under load-bearing conditions. As reviewed here, the
ARTICLE IN PRESS
Table 5
Overview of in vitro investigated biodegradable polymer/inorganic phase composites available in the literature
Composite PLGA/45S5 Bioglass
s
PDLLA/45S5 Bioglass
s
PLGA/HA PCL/HA
Cell culture Mouse broblasts (L929) [15],
marrow stromal cells [19],
mouse broblasts (208F) [86],
human osteosarcoma cells
(SaOS-2) [148]
Mouse broblasts (208F) [137],
human osteosarcoma cell line
(MG-63) [134], human lung
carcinoma (A549) [134,136]
Rat calvarial osteoblasts [194] Human osteosarcoma cells
(SaOS-2), osteoblasts from
human trabecular bone [195]
K. Rezwan et al. / Biomaterials 27 (2006) 34133431 3425
mechanical integrity of man-made composite scaffolds is
still at least one order of magnitude lower than that of
cancellous or cortical bone. Achieving the mechanical
properties of bone might also allow replacing bigger parts
of damaged bone tissue than what is possible today.
The incorporation of biomolecules such as growth
factors with the aim to accelerate local bone healing is
promising and currently under extensive research. Incor-
porating biomolecules during scaffold processing however
is not simple as biomolecules are sensitive to elevated
temperatures and extreme chemical conditions. A promis-
ing strategy is the immobilization of proteins and growth
factors in the post-processing phase via surface functiona-
lization of the scaffold.
Hardly any in vitro and particularly in vivo studies exist
for the composite scaffolds reviewed in this article.
However, in order to target clinical applications, in vitro
and in vivo studies are inevitable and the need for more
studies in biological systems is imperative.
6.2. Future trend: stem cells and ideal biomaterials
The application of biomaterials in hard tissue repair
started with bioinert approaches, which involved the
development and application of bioinert materials. These
materials are applied in most permanent bioimplants in
todays clinical use such as, for example, hipjoint
replacements. The subsequent development of biomaterials
focused on bone-bonding properties of bioactive glasses
and ceramics [81]. This period was soon followed by the
development of biodegradable materials for bone tissue
engineering scaffolds that can stimulate specic cellular
responses at the molecular level [23].
The composite scaffolds reviewed in this article combine
the features of the biomaterials in the second and third
periods: they possess bioactivity, degradability and the
possibility of biomolecule incorporation. Over the years,
the developed biomaterials have addressed biological
aspects of increased complexity, starting on the level of
ion interactions and moving then to growth factor
incorporation. The biomaterials were extended from purely
synthetic materials to material/biologic hybrids, engineer-
ing at the same time bioactivity and biodegradability.
While current research is still focused on the interaction
between stromal cells and biomaterials, the fundaments for
biomaterials seem to originate from introducing stem cells.
Scaffolds seeded with stem cells allow local cell function
adaptation by differentiation of stem cells as demonstrated
by Levenberg et al. in 2003 [8,196]. This new approach
enables the scaffold surface to mimic complex local
biological functions and may lead in near future to in
vitro and in vivo growth of tissues and organs. The
interface of stem cells and scaffolds are at the moment in
the center of attention, issuing growth factor incorporation
and cell adhesion [8]. In this approach, we anticipate that
engineered composite scaffolds made by biodegradable
polymer matrices with bioactive inorganic phases, as
reviewed here, will play a vital role and perhaps they will
be the scaffolds of choice in combination with stem cell
seeding.
Acknowledgment
Stimulating discussions with Prof. L.L. Hench, Prof. J.
Polak and Dr. A. Bishop (TERM-Centre, Imperial College
London) and with Dr. F. Filser (ETH Zurich) are
acknowledged.
References
[1] Williams D. Benet and risk in tissue engineering. Mater Today
2004;7:249.
[2] Mano JF, Sousa RA, Boesel LF, Neves NM, Reis RL. Bioinert,
biodegradable and injectable polymeric matrix composites for hard
tissue replacement: state of the art and recent developments.
Compos Sci Technol 2004;64:789817.
[3] Shin H, Jo S, Mikos AG. Biomimetic materials for tissue
engineering. Biomaterials 2003;24:435364.
[4] Drotleff S, Lungwitz U, Breunig M, Dennis A, Blunk T, Tessmar J.
Biomimetic polymers in pharmaceutical and biomedical sciences.
Eur J Pharm Biopharm 2004;58:385407.
[5] Suchanek W, Yoshimura M. Processing and properties of hydro-
xyapatite-based biomaterials for use as hard tissue replacement
implants. J Mater Res 1998;13:94117.
[6] Chaikof EL, Matthew H, Kohn J, Mikos AG, Prestwich GD, Yip
CM. Biomaterials and scaffolds in reparative medicine. Ann NY
Acad Sci 2002;961:96105.
[7] Antoniou G, Mikos AG, Temenoff JS. Formation of highly porous
biodegradable scaffolds for tissue engineering. Electron J Biotechnol
2000;3.
[8] Levenberg S, Langer R. Advances in tissue engineering current
topics in developmental biology, vol. 61. New York: Academic
Press; 2004 (p. 113134).
[9] Grifth LG. Emerging design principles in biomaterials and
scaffolds for tissue engineering. Ann NY Acad Sci 2002;961:8395.
[10] Karageorgiou V, Kaplan D. Porosity of 3D biomaterial scaffolds
and osteogenesis. Biomaterials 2005;26:547491.
[11] Tirelli N, Lutolf MP, Napoli A, Hubbell JA. Poly(ethylene glycol)
block copolymers. Rev Mol Biotechnol 2002;90:315.
[12] Berger J, Reist M, Mayer JM, Felt O, Peppas NA, Gurny R.
Structure and interactions in covalently and ionically crosslinked
chitosan hydrogels for biomedical applications. Eur J Pharm
Biopharm 2004;57(1):1934.
[13] Hench LL. Bioceramics. J Am Ceram Soc 1998;81(7):170528.
[14] Kim HW, Knowles JC, Kim HE. Hydroxyapatite/poly([epsilon]-
caprolactone) composite coatings on hydroxyapatite porous bone
scaffold for drug delivery. Biomaterials 2004;25:127987.
[15] Boccaccini AR, Blaker JJ, Maquet V, Day RM, Jerome R.
Preparation and characterisation of poly(lactide-co-glycolide)
(PLGA) and PLGA/Bioglass(R) composite tubular foam scaffolds
for tissue engineering applications. Mater Sci Eng C 2005;25:2331.
[16] Hedberg EL, Shih CK, Lemoine JJ, Timmer MD, Liebschner MAK,
Jansen JA. In vitro degradation of porous poly(propylene fuma-
rate)/poly(DL-lactic-co-glycolic acid) composite scaffolds. Biomater-
ials 2005;26:321525.
[17] Zhang K, Wang Y, Hillmayer MA, Francis LF. Processing and
properties of porous poly(L-lactide)/bioactive glass composites.
Biomaterials 2004;25:2489500.
[18] Niiranen H, Pyha lto T, Rokkanen P, Kelloma ki M, To rma la P. In
vitro and in vivo behavior of self-reinforced bioabsorbable polymer
and self-reinforced bioabsorbable polymer/bioactive glass compo-
sites. J Biomed Mater Res A 2004;69A:699708.
ARTICLE IN PRESS
K. Rezwan et al. / Biomaterials 27 (2006) 34133431 3426
[19] Yao J, Radin S, Leboy PS, Ducheyne P. The effect of bioactive glass
content on synthesis and bioactivity of composite poly (lactic-co-
glycolic acid)/bioactive glass substrate for tissue engineering.
Biomaterials 2005;26:193543.
[20] Gittens SA, Uludag H. Growth factor delivery for bone tissue
engineering. J Drug Target 2001;9:40729.
[21] Luginbuehl V, Meinel L, Merkle HP, Gander B. Localized delivery
of growth factors for bone repair. Eur J Pharm Biopharm
2004;58:197208.
[22] Di Silvio L, Boneld W. Biodegradable drug delivery system for the
treatment of bone infection and repair. J Mater SciMater Med
1999;10:6538.
[23] Hench LL, Polak JM. Third-generation biomedical materials.
Science 2002;295:10147.
[24] Maquet V, Boccaccini AR, Pravata L, Notingher I, Jerome R.
Porous poly([alpha]-hydroxyacid)/Bioglass(R) composite scaffolds
for bone tissue engineering. I: Preparation and in vitro characterisa-
tion. Biomaterials 2004;25:418594.
[25] Li H, Chang J. Preparation and characterisation of bioactive and
biodegradable wollastonite/poly(D,L-lactic acid) composite scaf-
folds. J Mater SciMater Med 2004;15:108995.
[26] Lu HH, Tang A, Oh SC, Spalazzi JP, Dionisio K. Compositional
effects on the formation of a calcium phosphate layer and the
response of osteoblast-like cells on polymer-bioactive glass compo-
sites. Biomaterials 2005;26:632334.
[27] Jiang G, Evans ME, Jones I, Rudd CD, Scotchford CA, Walker GS.
Preparation of poly(e-caprolactone)/continuous bioglass bre com-
posite using monomer transfer moulding for bone implant.
Biomaterials 2005;26:22818.
[28] Khan YM, Dhirendra DS, Katti S, Laurencin CT. Novel polymer-
synthesized ceramic composite-based system for bone repair: an in
vitro evaluation. J Biomed Mater Res A 2004;69A:72837.
[29] Xu HHK, Quinn JB, Takagi S, Chow LC. Synergistic reinforcement
of in situ hardening calcium phosphate composite scaffold for bone
tissue engineering. Biomaterials 2004;25:102937.
[30] Reis RL, Cunha AM, Allan PS, Bevis MJ. Mechanical behavior of
injection-molded starch-based polymers. Polym Adv Technol
1996;7:78490.
[31] Seal BL, Otero TC, Panitch A. Polymeric biomaterials for tissue and
organ regeneration. Mater Sci Eng: R: Rep 2001;34:147230.
[32] Di Martino A, Sittinger M, Risbud MV. Chitosan: a versatile
biopolymer for orthopaedic tissue-engineering. Biomaterials 2005;
26:598390.
[33] Lee SB, Kim YH, Chong MS, Hong SH, Lee YM. Study of gelatin-
containing articial skin V: fabrication of gelatin scaffolds using a
salt-leaching method. Biomaterials 2005;26:19618.
[34] Mohanty AK, Misra M, Hinrichsen G. Biobres, biodegradable
polymers and biocomposites: an overview. Macromol Mater Eng
2000;276277:124.
[35] Chen GQ, Wu Q. The application of polyhydroxyalkanoates as
tissue engineering materials. Biomaterials 2005;26:656578.
[36] Middleton JC, Tipton AJ. Synthetic biodegradable polymers as
orthopaedic devices. Biomaterials 2000;21:233546.
[37] Lu LC, Mikos AG. Poly(lactic acid). In: Mark JE, editor. Polymer
data handbook. Oxford: Oxford Press; 1999. p. 527633.
[38] Yang S, Leong KF, Du Z, Chua CK. The design of scaffolds for use
in tissue engineering. Part I. Traditional factors. Tissue Eng
2001;7:67989.
[39] Ramakrishna S, Huang ZM, Kumar GV, Batchelor AW, Mayer J.
An introduction to biocomposites. London: Imperial College Press;
2004 (p. 36).
[40] Gunatillak PA, Adhikari R. Biodegradable synthetic polymers for
tissue engineering. Eur Cells Mater J 2003;5:116.
[41] Iroh JO. Poly(epsilon-caprolactone). In: Mark JE, editor. Polymer
data handbook. Oxford: Oxford Press; 1999. p. 3612.
[42] Kellomak M, Heller J, Tormala P. Processing and properties of two
different poly(ortho esters). J Mater SciMater Med 2000;11:
34555.
[43] Magill JH. Poly(phosphazenes), Bioerodible. In: Mark JE,
editor. Polymer data handbook. Oxford: Oxford Press; 1999.
p. 7469.
[44] Kumudine C, Premachandra JK. Poly(lactic acid). In: Mark JE,
editor. Polymer data handbook. Oxford: Oxford Press; 1999.
p. 707.
[45] Kohn J, R L. Bioresorbable and bioerodible materials. In: Ratner
BD, Hoffman AS, Schoen FJ, JE L, editors. Biomaterials science: an
introduction to materials in medicine. New York: Academic Press;
1996. p. 6472.
[46] Jagur-Grodzinski J. Biomedical application of functional polymers.
Reactive Funct Polym 1999;39:99138.
[47] Bergsma EJ, Rozema FR, Bos RRM, Debruijn WC. Foreign body
reaction to resorbable poly(L-lactic) bone plates and screws used for
the xation of unstable zygomatic fractures. J Oral Maxillofac Surg
1993;51:66670.
[48] Martin C, Winet H, Bao JY. Acidity near eroding polylactide-
polyglycolide in vitro and in vivo in rabbit tibial bone chambers.
Biomaterials 1996;17(24):237380.
[49] Andrew SD, Phil GC, Marra KG. The inuence of polymer blend
composition on the degradation of polymer/hydroxyapatite bioma-
terials. J Mater Sci: Mater Med 2001;12:6737.
[50] Heidemann W, Jeschkeit S, Rufeux K, Fischer JH, Wagner M,
Kruger G, et al. Degradation of poly(D,L)lactide implants with or
without addition of calciumphosphates in vivo. Biomaterials
2001;22:237181.
[51] Rich J, Jaakkola T, Tirri T, Narhi T, Yli-Urpo A, Seppala J. In vitro
evaluation of poly([var epsilon]-caprolactone-co-DL-lactide)/bioac-
tive glass composites. Biomaterials 2002;23:214350.
[52] Boccaccini AR, Maquet V. Bioresorbable and bioactive polymer/
Bioglass(R) composites with tailored pore structure for tissue
engineering applications. Compos Sci Technol 2003;63:241729.
[53] Gollwitzer H, Thomas P, Diehl P, Steinhauser E, Summer B,
Barnstorf S, et al. Biomechanical and allergological characteristics
of a biodegradable poly(D,L-lactic acid) coating for orthopaedic
implants. J Orthop Res 2005;23:8029.
[54] Schmidmaier G, Wildemann B, Stemberger A, Haas NP, Raschke
M. Biodegradable poly(D,L-lactide) coating of implants for
continuous release of growth factors. J Biomed Mater Res 2001;
58:44955.
[55] Schmidmaier G, Wildemann B, Bail H, Lucke M, Fuchs T,
Stemberger A, et al. Local application of growth factors (insulin-
like growth factor-1 and transforming growth factor-[beta]1) from a
biodegradable poly(-lactide) coating of osteosynthetic implants
accelerates fracture healing in rats. Bone 2001;28:34150.
[56] Gollwitzer H, Ibrahim K, Meyer H, Mittelmeier W, Busch R,
Stemberger A. Antibacterial poly(D,L-lactic acid) coating of medical
implants using a biodegradable drug delivery technology.
J Antimicrob Chemother 2003;51:58591.
[57] Pitt CG, Gratzel MM, Kimmel GL. Aliphatic polyesters. 2. The
degradation of poly(DL-lactide), poly(e-caprolactone) and their
copolymers in vivo. Biomaterials 1981;2:21520.
[58] Gabelnick HL. Biodegradable implants: alternative approaches. In:
Mishell DR, editor. Advanced in human fertility and reproductive
endocrinology: vol. 2: Long acting steroid contraception.
New York: Raven Press; 1983. p. 14973.
[59] Payne RG, Mikos AG. Synthesis of synthetic polymers: poly
(propylene fumarate). In: Atala A, Lanza RP, editors. Methods of
tissue engineering. California: Academic Press; 2002. p. 64952.
[60] Yaszenski MJ, Payne RG, Hayes WC, Langer R, Aufdemorte TB,
Mikos AG. The ingrowth of new bone tissue and initial mechanical
properties of a degrading polymeric composite scaffold. Tissue Eng
1995;1:4152.
[61] Peter SJ, Lu L, Kim DJ, Stamatas GN, Miller MJ, Yaszemski MJ,
et al. Effects of transforming growth factor beta-1 released from
biodegradable polymer microparticles on marrow stromal osteo-
blasts cultured on poly(propylene fumarate) substrates. J Biomed
Mater Res 2000;50:45262.
ARTICLE IN PRESS
K. Rezwan et al. / Biomaterials 27 (2006) 34133431 3427
[62] Doi Y, Kitamura S, Abe H. Microbial synthesis and characteriza-
tion of poly(3-hydroxybutyrate-co-3-hydroxyhexanoate). Macromo-
lecules 1995;28:48228.
[63] Li HY, Du RL, Chang J. Fabrication, characterization, and in vitro
degradation of composite scaffolds based on PHBV and bioactive
glass. J Biomater Appl 2005;20:13755.
[64] Doyle C, Tanner ET, Boneld W. In vitro and in vivo evaluation of
polyhydroxybutyrate and of polyhydroxybutyrate reinforced with
hydroxyapatite. Biomaterials 1991;12:8417.
[65] Verma S, Bhatia Y, Valappil S, Roy I. A possible role of poly 3-
hydroxybutyric acid in antibiotic production in Streptomyces. Arch
Microbiol 2002;179:669.
[66] Shastri VP, Zelikin A, Hildgen P. Synthesis of synthetic polymers:
poly(anhydrides). In: Atala ALR, editor. Methods of tissue
engineering. California: Academic Press; 2002. p. 60917.
[67] Hench LL. Bioceramics: from concept to clinic. J Am Ceram Soc
1991;74:1487510.
[68] Kokubo T, Kim H-M, Kawashita M. Novel bioactive materials with
different mechanical properties. Biomaterials 2003;24:216175.
[69] Chen QZ, Boccaccini AR. Poly(D,L-lactic acid) coated 45S5
Bioglass-based scaffolds for bone engineering. J Biomed Mater
Res A, in press.
[70] LeGeros RZ, LeGeros JP. Dense hydroxyapatite. In: Hench LL,
Wilson J, editors. An introduction to bioceramics. 2nd ed. London:
Word Scientic; 1999. p. 13980.
[71] Hench LL. Bioactive glasses and glasses-ceramics. In: Shackelford
JF, editor. Bioceramics -applications of ceramic and glass materials
in medicine. Switzerland: Trans Tech Publication; 1999. p. 3764.
[72] Hench LL, Kokubo T. Properties of bioactive glasses and glass-
ceramics. In: Black J, Hastings G, editors. Handbook of biomaterial
properties. London: Chapman & Hall; 1998. p. 35563.
[73] Kokubo T. A/W glass-ceramic: processing and properties. In:
Hench LL, Wilson J, editors. An introduction to bioceramics. 2nd
ed. London: Word Scientic; 1999. p. 7588.
[74] Keaveny TM, Hayes WC. Mechanical properties of cortical and
trabecular bone. In: Hall BK, editor. Bone growth. Boca Raton, FL:
CRC Press; 1993. p. 285344.
[75] Zioupos P, Currey JD. Changes in the stiffness, strength,
and toughness of human cortical bone with age. Bone 1998;22:
5766.
[76] Hench LL, Wilson J, editors. An introduction to bioceramics. 2nd
ed. London: Word Scientic; 1999.
[77] Hench LL, Splinter RJ, Allen WC. Bonding mechanisms at the
interface of ceramic prosthetic materials. J Biomed Mater Res Symp
1971;2:11741.
[78] Clark AE, Hench LL. Calcium phosphate formation on solgel
derived bioactive glasses. J Biomed Mater Res 1994;28:6938.
[79] Hench LL. Solgel materials for bioceramic applications. Curr Opin
Solid State Mater Sci 1997;2:60410.
[80] Wilson J, Pigott GH, Schoen FJ, Hench LL. Toxicology and
biocompatibility of Bioglass. J Biomed Mater Res 1981;15:80511.
[81] Hench LL, Wilson J. Surface-active biomaterials. Science 1984;226:
6306.
[82] Lobel KD, Hench LL. In-vitro protein interactions with a bioactive
gel-glass. J SolGel Sci Technol 1996;7:6976.
[83] Lobel KD, Hench LL. In vitro adsorption and activity of enzymes
on reaction layers of bioactive glass substrates. J Biomed Mater Res
1998;39:5759.
[84] Ohgushi H, Dohi Y, Yoshikawa T, Tamai S, Tabata S, Okunaga K,
et al. Osteogenic differentiation of cultured marrow stromal stem
cells on the surface of bioactive glass ceramics. J Biomed Mater Res
1996;32:3418.
[85] Aksay IA, Weiner S. Biomaterials-Is this really a eld of research?
Curr Opin Solid State Mater Sci 1998;3:21920.
[86] Day RM, Boccaccini AR, Shurey S, Roether JA, Forbes A, Hench
LL, et al. Assessment of polyglycolic acid mesh and bioactive
glass for soft-tissue engineering scaffolds. Biomaterials 2004;25:
585766.
[87] Keshaw H, Forbes A, Day RM. Release of angiogenic growth
factors from cells encapsulated in alginate beads with bioactive
glass. Biomaterials 2005;26:41719.
[88] Schepers E, de Clercq M, Ducheyne P, Kempeneers R. Bioactive
glass particulate material as ller for bone lesions. J Oral Rehab
1991;18:43952.
[89] Gatti AM, Valdre G, OH A. Analysis of the in vivo reactions of a
bioactive glass in soft and hard tissue. Biomaterials 1994;15:20812.
[90] Roether JA, Gough JE, Boccaccini AR, Hench LL, Maquet V,
Jerome R. Novel bioresorbable and bioactive composites based on
bioactive glass and polylactide foams for bone tissue engineering.
J Mater SciMater Med 2002;13:120714.
[91] Xynos ID, Edgar AJ, Buttery LDK, Hench LL, Polak JM. Ionic
products of bioactive glass dissolution increase proliferation of
human osteoblasts and induce insulin-like growth factor II mRNA
expression and protein synthesis. Biochem Biophys Res Commun
2000;276:4615.
[92] Xynos ID, Hukkanen MVJ, Buttery LDK, Hench LL, Polak JM.
Bioglass
s
45S5 stimulates osteoblast turnover and enhances bone
formation in vitro: implications and applications for bone tissue
engineering. Calcif Tissue Int 2000;67:3219.
[93] Laurencin CT, Lu HH, Y K. Processing of polymer scaffolds:
polymerceramic composite foams. In: Atala A, Lanza RP, editors.
Methods of tissue engineering. California: Academic Press; 2002.
p. 70514.
[94] Livingston T, Ducheyne P, Garino J. In vivo evaluation of a
bioactive scaffold for bone tissue engineering. J Biomed Mater Res
2002;62:113.
[95] Kaufmann EABE, Ducheyne P, Shapiro IM. Evaluation of
osteoblast response to porous bioactive glass (45S5) substrates by
RT-PCR analysis. Tissue Eng 2004;6:1928.
[96] Yuan H, de Bruijn JD, Zhang X, van Blitterswijk CA, de Groot K.
Bone induction by porous glass ceramic made from Bioglass
s
(45S5). J Biomed Mater Res B: Appl Biomater 2001;58:2706.
[97] Jones JR, Hench LL. Regeneration of trabecular bone using porous
ceramics. Curr Opin Solid State Mater Sci 2003;7:3017.
[98] Boccaccini AR, Notingher I, Maquet V, Je ro me R. Bioresorbable
and bioactive composite materials based on polylactide foams lled
with and coated by Bioglass
s
particles for tissue engineering
applications. J Mater Sci: Mater Med 2003;14:44350.
[99] Jones JR, Hench LL. Factors affecting the structure and properties
of bioactive foam scaffolds for tissue engineering. J Biomed Mater
Res B: Appl Biomater 2003;68B:3644.
[100] Xynos ID, Edgar AJ, Buttery LDK, Hench LL, Polak M. Gene-
expression proling of human osteoblasts following treatment with
the ionic products of Bioglass 45S5 dissolution. J Biomed Mater Res
2001;55:1517.
[101] Bellantone M, Williams HD, Hench LL. Broad-spectrum bacter-
icidal activity of AgO2-doped bioactive glass. Antimicrob Agents
Chemother 2002;46:19405.
[102] Peitl O, LaTorre GP, Hench LL. Effect of crystallization on apatite-
layer formation of bioactive glass 45S5. J Biomed Mater Res 1996;
30:50914.
[103] Li P, Zhang F, Kokubo T. The effect of residual glassy phase in a
bioactive glass-ceramic on the formation of its surface apatite layer
in vitro. J Mater Sci: Mater Med 1992;3:4526.
[104] Clupper DC, Hench LL. Crystallization kinetics of tape cast
bioactive glass 45S5. J Non-Cryst Solids 2003;318:438.
[105] Jones JR, Ehrenfried LM, Hench LL. Optimising bioactive glass
scaffolds for bone tissue engineering. Biomaterials 2006;27:96473.
[106] Kim HW, Knowles JC, Kim HE. Hydroxyapatite porous scaffold
engineered with biological polymer hybrid coating for antibiotic
Vancomycin release. J Mater Sci: Mater Med 2005;16:18995.
[107] Giesen EBW, Ding M, Dalstra M, van Eijden TMGJ. Mechanical
properties of cancellous bone in the human mandibular condyle are
anisotropic. J Biomech 2001;34:799803.
[108] Yeni YN, Fyhrie DP. Finite element calculated uniaxial apparent
stiffness is a consistent predictor of uniaxial apparent strength in
ARTICLE IN PRESS
K. Rezwan et al. / Biomaterials 27 (2006) 34133431 3428
human vertebral cancellous bone tested with different boundary
conditions. J Biomech 2001;34:164954.
[109] Kokubo T. A/W glass-ceramic: processing and properties. In:
Hench LL, Wilson J, editors. An introduction to bioceramics. 2nd
ed. London: Word Scientic; 1999. p. 7588.
[110] LeGeros RZ, LeGeros JP. Calcium phosphate ceramics: pas, present
and future. Bioceramics, vol.15, Proceedings of the 15th interna-
tional symposium on ceramics in medicine, Sydney, 2002. p. 310.
[111] Yamamuro T, Hench LL, Wilson J. Handbook of bioactive
ceramics. Boca Raton, FL: CRC Press; 1990.
[112] Knowles JC. A review article: phosphate glasses for biomedical
applications. J Mater Chem 2003;13:2395401.
[113] Jarcho M, Kay JF, Gumar KI, Doremus RH, Drobeck HP. Tissue,
cellular and subcellular events at a bone-ceramic hydroxylapatite
interface. J Biosci Bioeng 1977;1:7992.
[114] Hammerle CH, Olah AJ, Schid J, Fluckiger L, Gogolewski S,
Winkler JR, et al. The biological effect of natural bone mineral on
bone neoformation on the rabbit skull. Clin Oral Implants Res
1997;8:198207.
[115] Denissen HW, deGroot K, Kakkes P, van den Hooff A, Klopper PJ.
Tissue response to dense apatite implants in rats. J Biomed Mater
Res 1980;14:71321.
[116] Hollinger JO, Battistone GC. Biodegradable bone repair materials.
Clin Orthop Rel Res 1986;207:290305.
[117] Brown S, Clarke I, Williams P. Bioceramics 14: Proceedings of the
14th international symposium on ceramics in medicine, Interna-
tional symposium on ceramics in medicine, International sympo-
sium on ceramics in medicine, Palm Springs, CA, 2001. p. 21369.
[118] Oonishi H, Kushitani S, Iwaki H. Comparative bone formation in
several kinds of bioceramic granules. In: Wilson J, Hench LL,
Greenspan D, editors. Eighth international symposium on ceramics
in medicine. Tokyo, Japan: Elsevier Science Ltd.; 1995. p. 13744.
[119] de Groot K, Lein CPAT, Wolke JGC, de Bliek-Hogervost JMA.
Chemistry of calcium phosphate bioceramics. In: Yamamuro T,
Hench LL, Wilson J, editors. Handbook of bioactive ceramics. Boca
Raton, FL: CRC Press; 1990. p. 316.
[120] Miyazaki T, Kim HM, Kokubo T, Kato H, Nakamura T. Induction
and acceleration of bonelike apatite formation on tantalum oxide
gel in simulated body uid. J SolGel Sci Technol 2001;21:838.
[121] Miyazaki T, Kim HM, Kokubo T, Ohtsuki C, Nakamura T.
Apatite-forming ability of niobium oxide gels in a simulated body
uid. J Ceram Soc Japan 2001;109:92933.
[122] Li P, Ohtsuki C, Kokubo T, Nakanishi K, Soga N, Nakamura T,
et al. Apatite formation induced by silica gel in a simulated body
uid. J Am Ceram Soc 1992;75:20947.
[123] Uchida M, Kim HM, Kokubo T, Miyaji F, Nakamura T. Bonelike
apatite formation induced on zirconia gel in a simulated body uid
and its modied solutions. J Am Ceram Soc 1992;84:20414.
[124] Li PJ, Ohtsuki C, Kokubo T, Nakanishi K, Soga N, deGroot K.
The role of hydrated silica, titania, and alumina in inducing apatite
on implants. J Biomed Mater Res 1994;28:715.
[125] Ramakrishna S, Mayer J, Wintermantel E, Leong KW. Biomedical
applications of polymer-composite materials: a review. Compos Sci
Technol 2001;61:1189224.
[126] Wang M. Developing bioactive composite materials for tissue
replacement. Biomaterials 2003;24:213351.
[127] Zhang RY, Ma PX. Poly(alpha-hydroxyl acids)/hydroxyapatite
porous composites for bone-tissue engineering. I. Preparation and
morphology. J Biomed Mater Res 1999;44:44655.
[128] Verheyen CCPM, deWijn J, van Blitterswijk C, deGroot K, Rozing
P. Hydroxyapatite/poly(L-lactide) composites: an animal study on
push-out strengths and interface histology. J Biomed Mater Res
1993;27:43344.
[129] Durucan C, Brown PW. Biodegradable hydroxyapatite-polymer
composites. Adv Eng Mater 2001;3:22731.
[130] Korventausta J, Jokinen M, Rosling A, Peltola T, Yli-Urpo A.
Calcium phosphate formation and ion dissolution rates from silica
gel-PDLLA composites. Biomaterials 2003;24:517382.
[131] Boccaccini AR, Blaker JJ, Maquet V, Chung W, Jerome R, Nazhat
SN. PDLLA foams with TiO2 nanoparticles and PDLLA/TiO2-
Bioglass foam composites for tissue engineering scaffolds. J Mater
Sci, in press.
[132] Blaker JJ, Maquet V, Jerome R, Boccaccini AR, SN Nazhat.
Mechanically anisotropic PDLLA/Bioglass composite foams as
scaffolds for bone tissue engineering. Acta Biomater 2005;1:64352.
[133] Maquet V, Boccaccini AR, Pravata L, Notingher I, Je ro me R.
Preparation, characterization, and in vitro degradation of bior-
esorbable and bioactive composites based on Bioglass
s
-lled
polylactide foams. J Biomed Mater Res A 2003;66A:33546.
[134] Blaker JJ, Gough JE, Maquet V, Notingher I, Boccaccini AR. In
vitro evaluation of novel bioactive composites based on Bioglass
s
-
lled polylactide foams for bone tissue engineering scaffolds. J
Biomed Mater Res A 2003;67A:140111.
[135] Roether JA, Boccaccini AR, Hench LL, Maquet V, Gautier S,
Jerome R. Development and in vitro characterisation of novel
bioresorbable and bioactive composite materials based on polylac-
tide foams and Bioglass(R) for tissue engineering applications.
Biomaterials 2002;23:38718.
[136] Verrier S, Blaker JJ, Maquet V, Hench LL, Boccaccini AR.
PDLLA/Bioglass(R) composites for soft-tissue and hard-tissue
engineering: an in vitro cell biology assessment. Biomaterials
2004;25:301321.
[137] Day RM, Maquet V, Boccaccini AR, Je ro me R, Forbes A. In vitro
and in vivo analysis of macroporous biodegradable poly(d,l-lactide-
co-glycolide) scaffolds containing bioactive glass. J Biomed Mater
Res 2005;75A:77887.
[138] Jaakkola T, Rich J, Tirri T, Narhi T, Jokinen M, Seppala J, et al. In
vitro Ca-P precipitation on biodegradable thermoplastic composite
of poly([epsilon]-caprolactone-co-lactide) and bioactive glass
(S53P4). Biomaterials 2004;25:57581.
[139] Li HY, Chang J. pH-compensation effect of bioactive inorganic
llers on the degradation of PLGA. Compos Sci Technol
2005;65:222632.
[140] Shikinami Y, Okuno M. Bioresorbable devices made of forged
composites of hydroxyapatite (HA) particles and poly -lactide
(PLLA). Part II: practical properties of miniscrews and miniplates.
Biomaterials 2001;22:3197211.
[141] Sepulveda P, Jones JR, Hench LL. Bioactive solgel foams for tissue
repair. J Biomed Mater Res 2001;59:3408.
[142] Lee LJ, Zeng C, Cao X, Han X, Shen J, Xu G. Polymer
nanocomposite foams. Compos Sci Technol 2005;65:234463.
[143] Ma PX, Zhang R. Microtubular architecture of biodegradable
polymer scaffolds. J Biomed Mater Res 2001;56:46977.
[144] Thomson RC, Yaszemski MJ, Powers JM, Mikos AG. Hydro-
xyapatite bber reinforced poly( -hydroxy ester) foams for bone
regeneration. Biomaterials 1998;12:193543.
[145] Aho AJ, Tirri T, Kukkonen J. Injectable bioactive glass/biodegrad-
able polymer composite for bone and cartilage reconstruction:
Concept and experimental outcome with thermoplastic composites
of poly(epsilon-caprolactone-co-D,L-lactide) and bioactive glass
S53P4. J Mater SciMater Med 2004;15:116573.
[146] Xiong Z, Yan YN, Wang SG, Zhang RJ, Zhang C. Fabrication of
porous scaffolds for bone tissue engineering via low-temperature
deposition. Scr Mater 2002;46:7716.
[147] Taboas JM, Maddox RD, Krebsbach PH, Hollister SJ. Indirect
solid free form fabrication of local and global porous, biomimetic
and composite 3D polymer-ceramic scaffolds. Biomaterials 2003;
24:18194.
[148] Lu HH, El-Amin SF, Scott KD, Laurencin CT. Three-dimensional,
bioactive, biodegradable, polymer-bioactive glass composite scaf-
folds with improved mechanical properties support collagen
synthesis and mineralization of human osteoblast-like cells in vitro.
J Biomed Mater Res A 2003;64A:46574.
[149] Deng X, Hao J, Wang C. Preparation and mechanical properties of
nanocomposites of poly(D,L-lactide) with Ca-decient hydroxyapa-
tite nanocrystals. Biomaterials 2001;22:286773.
ARTICLE IN PRESS
K. Rezwan et al. / Biomaterials 27 (2006) 34133431 3429
[150] Kasaga T, Yoshio O, Masayuki N, Yoshihiro A. Preparation and
mechanical properties of polylactide acid composites containing
hydroxyapatite bres. Biomaterials 2001;22:923.
[151] Xu HHK, Simon JCG. Self-hardening calcium phosphate com-
posite scaffold for bone tissue engineering. J Orthop Res 2004;22:
53543.
[152] Kikuchi M, Tanaka J, Koyama Y, Takakuda K. Cell culture tests of
TCP/CPLA composite. J Biomed Mater Res 1999;48:10810.
[153] Peter SJ, Miller ST, Zhu G, Yasko AW, Mikos AG. In vivo
degradation of poly(propylene fumarate)/beta-tricalcium phos-
phate injectable composite scaffold. J Biomed Mater Res 1998;41:
17.
[154] Juhasz JA, Best SM, Brooks R, Kawashita M, Miyata N, Kokubo
T, et al. Mechanical properties of glass-ceramic A-W-polyethylene
composites: effect of ller content and particle size. Biomaterials
2004;25:94955.
[155] Ambrosio AMA, Sahota JS, Khan Y, Laurencin CT. A novel
amorphous calcium phosphate polymer ceramic for bone repair: I.
synthesis and characterization. J Biomed Mater Res 2001;58:
295301.
[156] Guan L, Davies JE. Preparation and characterisation of a highly
macroporous biodegradable composite tissue engineering scaffold.
J Biomed Mater Res 2004;71A:4807.
[157] Devin JE, Attawia MA, Laurencin CT. Three-dimensional degrad-
able porous polymer-ceramic matrices for use in bone repair.
J Biomater Sci Polym Ed 1996;7:6619.
[158] Stamboulis AG, Boccaccini AR, Hench LL. Novel biodegradable
polymer/bioactive glass composites for tissue engineering applica-
tions. Adv Eng Mater 2002;4:1059.
[159] Navarro M, Ginebra MP, Planell JA, Zeppetelli S, Ambrosio L.
Development and cell response of a new biodegradable composite
scaffold for guided bone regeneration. J Mater Sci: Mater Med
2004;15:41922.
[160] Boccaccini AR, Blaker JJ. Bioactive composite materials
for tissue engineering scaffolds. Expert Opin Med Dev 2005;2:
30317.
[161] Weng J, Wang M, Chen J. Plasma-sprayed calcium phosphate
particles with high bioactivity and their use in bioactive scaffolds.
Biomaterials 2002;23:26239.
[162] Gibson LJ. The mechanical behaviour of cancellous bone.
J Biomech 1985;18:31728.
[163] Maquet V, Martin D, Scholtes F, Franzen R, Schoenen J, Moonen
G, et al. Poly(D,L-lactide) foams modied by poly(ethylene oxide)-
block-poly(D,L-lactide) copolymers and a-FGF: in vitro and in vivo
evaluation for spinal cord regeneration. Biomaterials 2001;22:
113746.
[164] Blacher S, Maquet V, Jerome R, Pirard JP, Boccaccini AR. Study of
the connectivity properties of Bioglass(R)-lled polylactide foam
scaffolds by image analysis and impedance spectroscopy. Acta
Biomater 2005;1:56574.
[165] Lam KH, Nieuwenhuis P, Molenaar I. Biodegradation of porous
versus non-porous poly(L-lactic acid) lms. J Mater Sci: Mater Med
1994;5:1819.
[166] Day RM, Boccaccini AR, Maquet V, Shurey S, Forbes A, Gabe
SM, et al. In vivo characterization of a novel bioresorbable poly
(lactide-co-glycolide) tubular foam scaffold for tissue engineering
applications. J Mater Sci: Mater Med 2004;15:72934.
[167] Gutwein LG, Webster T. Osteoblast and chrondrocyte proliferation
in the presence of alumina and titania nanoparticles. J Nanoparti-
culate Res 2002;4:2318.
[168] Schiller C, Siedler M, Peters F, Epple M. Functionally graded
materials of biodegradable polyesters and bone-like calcium
phosphates for bone replacement. In: Trumble K, Bowman K,
Reimanis I, Sampath S, editors. Functionally graded materials.
Westerville, OH: The American Ceramic Society; 2000.
p. 97108.
[169] Hutmacher DW. Scaffolds in tissue engineering bone and cartilage.
Biomaterials 2001;21:252943.
[170] Xiong Z, Yan YN, Wang SG, Zhang RJ, C Z. Fabrication of
porous scaffolds for bone tissue engineering via low-temperature
deposition. Scr Mater 2002;46:7716.
[171] Boccaccini AR, Stamboulis AG, Rashid A, Roether J. Composite
surgical sutures with bioactive glass coating. J Biomed Mater Res B:
Appl Biomater 2003;67B:61826.
[172] Boccaccini AR, Zhitomirsky I. Application of electrophoretic and
electrolytic deposition techniques in ceramics processing. Curr Opin
Solid State Mater Sci 2002;6:25160.
[173] Olsen-Claire J, Boccaccini AR. Bioglass
s
coatings on biodegradable
poly(3-hydroxybutyrate) (P3HB) meshes for tissue engineering
scaffolds, Matwiss. u. Werkstofftech. 2006, in press.
[174] Oliveira AL, Malafaya PB, Reis RL. Sodium silicate gel as a
precursor for the in vitro nucleation and growth of a bone-like
apatite coating in compact and porous polymeric structures.
Biomaterials 2003;24:257584.
[175] Azevedo HS, Leonor IB, Alves CM, Reis RL. Incorporation
of proteins and enzymes at different stages of the preparation
of calcium phosphate coatings on a degradable substrate
by a biomimetic methodology. Mater Sci Eng C 2005;25:
16979.
[176] Taddei P, Tinti A, Reggiani M, Fagnano C. In vitro mineralization
of bioresorbable poly(epsilon-caprolactone)/apatite composites for
bone tissue engineering: a vibrational and thermal investigation.
J Mol Struct 2005;744:13543.
[177] Mano JF, Sousa RA, Boesel LF, Neves NM, Reis RL. Bioinert,
biodegradable and injectable polymeric matrix composites for hard
tissue replacement: state of the art and recent developments.
Compos Sci Technol 2004;64:789817.
[178] Koegler WS, Grifth LG. Osteoblast response to PLGA tissue
engineering scaffolds with PEO modied surface chemistries and
demonstration of patterned cell response. Biomaterials 2004;25:
281930.
[179] Ladron deGuevara-Fernandez S, Ragel CV, Vallet-Regi M.
Bioactive glass-polymer materials for controlled release of ibupro-
fen. Biomaterials 2003;24:403743.
[180] Murphy WL, Peters MC, Kohn DH, Mooney DJ. Sustained release
of vascular endothelial growth factor from mineralized poly(lactide-
co-glycolide) scaffolds for tissue engineering. Biomaterials 2000;21:
25217.
[181] Laurencin CT, Attawia MA, Lu LQ, Borden MD, Lu HH, Gorum
WJ, et al. Poly(lactide-co-glycolide)/hydroxyapatite delivery of
BMP-2-producing cells: a regional gene therapy approach to bone
regeneration. Biomaterials 2001;22:12717.
[182] Silva GA, Costa FJ, Coutinho OP, Radin S, Ducheyne P, Reis RL.
Synthesis and evaluation of novel bioactive composite starch/
bioactive glass microparticles. J Biomed Mater Res A 2004;70A:
4429.
[183] Anselme K. Osteoblast adhesion on biomaterials. Biomaterials
2000;21:66781.
[184] Mrksich M. What can surface chemistry do for cell biology? Curr
Opin Chem Biol 2002;6:7947.
[185] Ito A, Ino K, Kobayashi T, Honda H. The effect of RGD peptide-
conjugated magnetite cationic liposomes on cell growth and cell
sheet harvesting. Biomaterials 2005;26:618593.
[186] Yang F, Williams CG, Wang DA, Lee H, Manson PN, Elisseeff J.
The effect of incorporating RGD adhesive peptide in polyethylene
glycol diacrylate hydrogel on osteogenesis of bone marrow stromal
cells. Biomaterials 2005;26(10):59918.
[187] Kang IK, Kwon BK, Lee JH, Lee HB. Immobilization of
proteins on poly(methyl methacrylate) lms. Biomaterials 1993;14:
78792.
[188] Williams RA, Blanch HW. Covalent immobilization of protein
monolayers for biosensor applications. Biosens Bioelectron 1994;9:
15967.
[189] Heule M, Rezwan K, Cavalli L, Gauckler LJ. A miniaturized
enzyme reactor based on hierarchically shaped porous ceramic
microstruts. Adv Mater 2003;15:11914.
ARTICLE IN PRESS
K. Rezwan et al. / Biomaterials 27 (2006) 34133431 3430
[190] Linkhart TA, Mohan S, Baylink DJ. Growth factors for bone
growth and repair: IGF, TGF[beta] and BMP. Bone 1996;19(1,
Suppl. 1):S1S12.
[191] Jansen JA, Vehof JWM, Ruhe PQ, Kroeze-Deutman H, Kuboki Y,
Takita H, et al. Growth factor-loaded scaffolds for bone engineer-
ing. J Control Rel 2005;101:12736.
[192] Babensee JE, McIntire LV, Mikos AG. Growth factor delivery for
tissue engineering. Pharm Res 2000;17:497504.
[193] Darling AL, Sun W. 3D Microtomographic characterization of
precision extruded poly e-caprolactone scaffolds. J Mater Res B:
Appl Biomater 2004;70B:3117.
[194] Jung Y, Kim SS, Kim YH, Kim SH, Kim BS, Kim S, et al. A
poly(lactic acid)/calcium metaphosphate composite for bone tissue
engineering. Biomaterials 2005;26:631422.
[195] Causa F, Netti PA, Ambrosio L, Ciapetti G, Baldini N, Pagani S,
et al. Poly epsilon-caprolactone/hydroxyapatite composites for bone
regeneration: in vitro characterization and human osteoblast
response. J Biomed Mater Res A 2006;76A:15162.
[196] Levenberg S, Huang NF, Lavik E, Rogers AB, Itskovitz-Eldor J,
Langer R. Differentiation of human embryonic stem cells on three-
dimensional polymer scaffolds. Proc Natl Acad Sci USA 2003;
100:127416.
ARTICLE IN PRESS
K. Rezwan et al. / Biomaterials 27 (2006) 34133431 3431

Vous aimerez peut-être aussi